Vous êtes sur la page 1sur 145

Institut National Polytechnique de Toulouse (INP Toulouse)

Anthony Ruiz
9 Fvrier 2012

Simulations Numriques Instationnaires de la combustion turbulente et


transcritique dans les moteurs cryotechniques
(Unsteady Numerical Simulations of Transcritical Turbulent Combustion
in Liquid Rocket Engines)

cole doctorale et discipline ou spcialit :


ED MEGEP : nergtique et transferts

CERFACS

B. Cuenot - Chef de projet CERFACS, HDR


L. Selle - Charg de recherche CNRS

F. Dupoirieux - Matre de Recherches lONERA, HDR


J. Rveillon - Professeur l'Universit de Rouen

S. Candel - Professeur l'Ecole Centrale Paris


P. Chassaing - Professeur mrite l'INPT
D. Saucereau - Ingnieur SNECMA
Contents

1 Introduction 1
1.1 Operating principle of Liquid Rocket Engines . . . . . . . . . . . . . . . . 1
1.2 Combustion in LREs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Preliminary Definitions . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Experimental studies . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.3 Numerical studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 Study Plan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2 Governing Equations, Thermodynamics and Numerics 19


2.1 Navier-Stokes Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.1 Species diusion flux . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.2 Viscous stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.3 Heat flux vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.4 Transport coecients . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.2 Filtered Equations for LES . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.1 The filtered viscous terms . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.2 Subgrid-scale turbulent terms for LES . . . . . . . . . . . . . . . . 26
2.3 Models for the subgrid-stress tensor . . . . . . . . . . . . . . . . . . . . . . 28
2.3.1 Smagorinsky model . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3.2 WALE model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 Real-Gas Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4.1 Generalized Cubic Equation of State . . . . . . . . . . . . . . . . . 29
2.4.2 Primitive to conservative variables . . . . . . . . . . . . . . . . . . 29
2.5 CPU cost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3 A DNS study of turbulent mixing and combustion in the near-injector


region of Liquid Rocket Engines 33
3.1 Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.1.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.1.3 Characteristic Numbers and Reference Scales . . . . . . . . . . . . . 37
3.1.4 Computational Grid . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1.5 Numerical Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1.6 Initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.7 Chemical kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2 Cold Flow Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.1 Vortex Shedding . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.2 Comb-like structures . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.2.3 Scalar Dissipation Rate and turbulent mixing . . . . . . . . . . . . 54
3.2.4 Mean flow field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4 Contents

3.2.5 Influence of mesh resolution . . . . . . . . . . . . . . . . . . . . . . 60


3.3 Reacting Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3.1 Flame stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3.2 Vortex shedding and comb-like structures . . . . . . . . . . . . . . . 64
3.3.3 Combustion Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.3.4 Mean flow field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.3.5 Comparison of numerical results with existing experimental data . . 73
3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4 Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame Issuing from


a Coaxial Injector, with and without Inner Recess 77
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.2 Configuration and operating point . . . . . . . . . . . . . . . . . . . . . . . 79
4.2.1 The Mascotte facility . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.2.2 C60 operating point . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.2.3 Characteristic numbers and scales . . . . . . . . . . . . . . . . . . . 79
4.3 Numerical setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
4.3.1 Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
4.3.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 84
4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.4.1 Reference solution without recess . . . . . . . . . . . . . . . . . . . 87
4.4.2 Eects of recess . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

5 General Conclusions 105

A Thermodynamic derivatives 107


A.1 Getting the density from (P,T,Yi ) . . . . . . . . . . . . . . . . . . . . . . . 107
A.1.1 From the EOS to the cubic polynomial . . . . . . . . . . . . . . . . 107
A.1.2 The Cardan method . . . . . . . . . . . . . . . . . . . . . . . . . . 107

Bibliography 113
List of Figures

1.1 The main components of the Ariane 5 european space launcher. . . . . . . 1


1.2 Components of a booster. . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Operating principle of the Vulcain 2 engine [Snecma 2011]. . . . . . . . . . 3
1.4 Injection plate of the Vulcain 2 LRE, composed of 566 coaxial injectors [As-
trium 2011] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.5 Heat capacity isocontours computed with the Soave-Redlich-Kwong equa-
tion of state (see Eq. 2.44): white = 103 J/K/kg; black=104 J/K/kg. . . . 5
1.6 Shadowgraphs at successive instants (time between frames is 0.1 ms) of a
transcritical N2 /supercritical He mixing layer [Teshome et al. 2011]. . . . . 8
1.7 Flame shape visualizations using OH , in a coaxial LOx/GH2 injector [Ju-
niper 2001]: (a) instantaneous image, (b) top: time-averaged image; bot-
tom: Abel-transformed image. The pressure is 7 MPa. . . . . . . . . . . . 9
1.8 OH PLIF image of a LOx/GH2 cryogenic flame [Singla et al. 2007]. . . . . 10
1.9 Shadowgraphs of a transcritical H2 /O2 reacting flow at 6 MPa, taken at
successive instants (time between frames is 0.25 ms) [Locke et al. 2010]. . . 10
1.10 Schematic of coaxial jet injector and the near-field mixing layers [Schu-
maker & Driscoll 2009] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.11 Radial distributions of normalized density at dierent axial locations (T =
300 K, uinj = 15 m/s, Tinj = 120 K, Dinj = 254 m) [Zong & Yang 2006] . 14
1.12 Contours of temperature for the near-field region for (a) supercritical and
(b) transcritical mixing [Oefelein & Yang 1998]. . . . . . . . . . . . . . . . 15
1.13 LES computations of transcritical jet flames: (a) T = 1000 K iso-surface
colored by axial velocity in a reacting transcritical LOx/GH2 flow [Mat-
suyama et al. 2010]. (b) Visualization of a transcritical LOx/GCH4 flame:
(top) direct visualization from experiment [Singla 2005]; (bottom) T =
(T max + T min )/2 isosurface from LES [Schmitt et al. 2010a]. . . . . . . . . 17
1.14 Schematic of a LRE combustion chamber, showing the dierent levels of
study considered in the present thesis work. . . . . . . . . . . . . . . . . . 18

2.1 Transport coecients for O2 at 100 bar, showing the liquid-like to gas-like
transition of thermo-physical properties . a) Dynamic viscosity b) Thermal
conductivity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2 Evolution of the Lewis number and normalized heat capacity with temper-
ature. Cp,0 = 1700 J/K/kg. The peak of Cp at the pseudo-boiling point
creates a local minimum in the Lewis number. . . . . . . . . . . . . . . . . 24
2.3 Density of oxygen as a function of temperature, at 10 MPa, computed
with the Peng-Robinson and the Soave-Redlich-Kwong equations of state,
and compared to the NIST database [Lemmon et al. 2009]. The circle at
T = 172 K shows the pseudo-boiling point. . . . . . . . . . . . . . . . . . . 30
6 List of Figures

3.1 a) Typical coaxial injector of a LRE. b) Boundary conditions for the 2D


computational domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.2 Probes located in the mixing layer. . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Transverse cut through the initial solution of the cold flow, downstream
the lip. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.4 Temporal evolution of the minimum, mean and maximum pressures in the
computational domain, after instantaneous ignition of the cold flow case. . 41
3.5 Strained diusion flame computed in AVBP: streamlines superimposed on
the temperature field. Thermodynamic conditions correspond to the split-
ter case: hydrogen at 150 K from the top, oxygen at 100 K from the bottom
and ambient pressure is 10 MPa. . . . . . . . . . . . . . . . . . . . . . . . 42
3.6 Comparison of flame structure versus mixture fraction between AVBP and
CANTERA. (a) temperature and (b) HO2 mass fraction. The vertical bar
indicates the stoichiometric mixture fraction. . . . . . . . . . . . . . . . . . 43
3.7 Major species mass fractions as a function of mixture fraction for the non-
premixed counterflow flame configuration (lines) and equilibrium (symbols) 44
3.8 Non-reacting flow: axial velocity field, showing the shear-induced Kelvin-
Helmholtz instability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.9 Non-reacting flow: oxygen mass fraction field at the H2 corner of the lip,
at successive instants. The time interval between frames is 10 s . . . . . . 46
3.10 Non-reacting flow: O2 mass fraction field, showing the rapid mixing of the
two streams by a large range of vortical structures. . . . . . . . . . . . . . 46
3.11 Non-reacting flow: temporal evolution of the transverse velocity field. The
time interval between two snapshots is conv . . . . . . . . . . . . . . . . . 48
3.12 Non-reacting flow: temporal evolution of the oxygen mass-fraction field.
The time interval between two snapshots is conv . . . . . . . . . . . . . . . 49
3.13 Non-reacting flow: transverse velocity signal analysis. Top) temporal
signal; dominant harmonic (St = 0.14). Bottom) Power Spectrum
Density without spectrum averaging. . . . . . . . . . . . . . . . . . . . . . 50
3.14 Non-reacting flow: eect of the Welch averaging procedure on the spectrum
frequency resolution. Number of windows=2,4,8. . . . . . . . . . . . . . . . 50
3.15 Non-reacting flow: spatial evolution of the transverse velocity spectrum in
the wake of the lip. Eight Welch averaging windows are used. . . . . . . . 51
3.16 Non-reacting flow: power spectrum density of the squared transverse ve-
locity at i=10, j=12. Eight Welch averaging windows are used. . . . . . . . 52
3.17 Non-reacting flow: vorticity field superimposed on the high-density region
(fluid regions with a density that is higher than 0.5 = 0.5 (inj inj
H2 + O2 ) are
painted in black.). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.18 Non-reacting flow: O2 mass fraction, along with a grey isocontour of mid-
density ( = 637 kg.m3 ) and a white isocontour of high scalar dissipation
rate ( = 5 103 s1 ). The dash circle shows a region of low scalar dissipation
rate where mixing has already occurred. . . . . . . . . . . . . . . . . . . . 54
3.19 Non-reacting flow: conditioned average and maximum value of scalar dis-
sipation rate as a function of mixture fraction. . . . . . . . . . . . . . . . . 55
List of Figures 7

3.20 Non-reacting flow: average density field . . . . . . . . . . . . . . . . . . . . 56


3.21 Non-reacting flow: spreading angles , . . . . . . . . . . . . . . . . . . . . 57
3.22 Non-reacting flow: transverse cuts of the density at 4 axial positions be-
tween x=0h and x=6h. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.23 Non-reacting flow: transverse cuts of the RMS axial velocity normalized
by the velocity dierence Us = UHinj2 UOinj2
at x=1h and x=5h. . . . . . . . 58
3.24 Non-reacting flow: transverse cuts of the RMS transverse velocity normal-
ized by the velocity dierence Us = UHinj2 UOinj2
at x=1h and x=5h. . . . . 59
3.25 Instantaneous snapshots of the transverse velocity and the density for the
(a) h100 and (b) h500 meshes. . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.26 Mean profiles, at x = 5h for the h100 and h500 meshes of (a) axial velocity,
(b) transverse velocity and (c) O2 mass fraction . . . . . . . . . . . . . . . 61
3.27 RMS profiles, at x = 5h for the h100 and h500 meshes, of (a) axial velocity,
(b) transverse velocity and (c) O2 mass fraction. . . . . . . . . . . . . . . . 62
3.28 Temperature field with superimposed density gradient (green) and heat-
release (black: max heat release rate of case AVBP_RG (=1013 W/m3 );
grey: 10 % of case AVBP_RG (=1012 W/m3 ). . . . . . . . . . . . . . . . . 63
3.29 Close-up view of the flame stabilization zone behind the splitter. Tempera-
ture field with superimposed iso-contours of density gradient (green=4 107 kg/m4 )
and heat-release rate (grey=1012 W/m3 , black=1013 W/m3 ). . . . . . . . . 64
3.30 The flame/vortex interaction separates the flame into 2 zones, a near-
injector steady diusion flame between O2 and a mixture of H2 and H2 O,
and a turbulent flame developing further downstream. . . . . . . . . . . . . 65
3.31 Comparison of the density fields between a) the non-reacting flow and b)
the reacting flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.32 Curvature PDF of the median density (0.5 iso-contour for the cold and
reacting flows. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.33 Reacting flow: instantaneous fields of (a) hydrogen (b) oxygen and (c)
water mass fraction, with superimposed stoichiometric mixture fraction
isoline (black). The dash line in (b) shows the location of the 1D cut that
is used in Fig. 3.34. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.34 Cut through a pocket of oxygen diluted with combustion products. The
location of the 1D cut is shown by a dash line in Fig. 3.33(b). . . . . . . . 69
3.35 Flame structure in the mixture fraction space. counterflow diusion
flame at a = 3800 s1 , scatter plot of the present turbulent flame. . . . . 70
3.36 Average density field for the reacting flow. White iso-contour: = 0.9 in O2 . 71
3.37 Average fields for the reacting flow. (a) Temperature. (b) Heat release rate.
Black iso-contour: 0.9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.38 OH mass-fraction fields at four instants. Liquid oxygen and gaseous hy-
drogen are injected below and above the step, respectively. . . . . . . . . . 73
3.39 OH-PLIF images of the flame holding region. Liquid oxygen and gaseous
hydrogen are injected above and below the step, respectively (opposite
arrangement of Fig. 3.38). The OH distribution in the flame edge is shown
on the standard color scale. p = 6.3 MPa (fig. 5 of [Singla et al. 2007]) . . 74
8 List of Figures

3.40 Comparison between experimental signal of OH PLIF [Singla et al. 2007] at


transcritical conditions and time averaged heat release rate in the current
simulation (the numerical visualization is upside down to compare with
experimental visualization) . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

4.1 Square combustion chamber with optical access used in the Mascotte facility. 79
4.2 Geometry of the coaxial injectors: a) dimensions and b) recess length. . . . 80
4.3 Cut of the h5 mesh, colored by = V 1/3 , with V the local volume of cells. 81
4.4 Species mass fraction as a function of mixture fraction, in the infinitely fast
chemistry limit (Burke-Schumann solution). . . . . . . . . . . . . . . . . . 83
4.5 Turbulence injection profiles: a) mean axial velocity and b) Root Mean
Square velocity fluctuations. The superscript c refers to the centerline value. 85
4.6 Instantaneous visualization of the reacting flow: a) temperature and b)
axial velocity fields. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.7 Qualitative comparison between instantaneous experimental ((a) and (c))
and numerical ((b) and (d)) visualizations of the reacting flow: (a) OH
emission (b) T = 1500 K iso-surface (c) shadowgraph (d) mixture fraction
between 0 (black) and 1 (white). . . . . . . . . . . . . . . . . . . . . . . . . 89
4.8 Instantaneous a) axial velocity and b) temperature cuts. . . . . . . . . . . 90
4.9 Qualitative comparison between instantaneous experimental ((a) and (c))
and numerical ((b) and (d)) visualizations of the near-injector region of the
reacting flow: (a) OH emission (b) T = 1500 K iso-surface (c) shadow-
graph (d) mixture fraction (black=0; white=1). . . . . . . . . . . . . . . . 91
4.10 Cut of the average velocity magnitude (white = 0 m/s; black = 200 m/s)
with superimposed streamlines in black and density iso-contours in grey
(=100, 500, 1000 kg/m3 ), obtained with the h5 mesh. . . . . . . . . . . . 92
4.11 Time-averaged flame shapes, shown by (a) Abel-transformed OH emis-
sion, (b) mean temperature field with the h5 mesh, (c) mean temperature
field with the h10 mesh (white=80 K; black=3200 K). . . . . . . . . . . . . 94
4.12 Cut of the RMS fluctuations of the axial velocity. . . . . . . . . . . . . . . 95
4.13 NR configuration: comparison between H2 CARS measurements and con-
ditional temperatures T , defined in Eq. 4.11. : CARS measurements
from [Habiballah et al. 2006] and [Zurbach 2006]. (a) positions of the cuts
(b) y = 4 mm, (c) x = 15 mm, (d) x = 50 mm, (e) x = 100 mm. . . . . . . 97
4.14 Instantaneous visualization of the near-injector region in the R configura-
tion: (a) iso-surface of temperature (=1500 K) and (b) mixture fraction
field. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.15 Qualitative comparison for the R configuration, between an instantaneous
(a) experimental shadowgraph and (b) a mixture fraction field between 0
(black) and 1 (white). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.16 Time-averaged results. NR configuration (left) and R configuration (right):
(a) and (b) are experimental shadowgraphs and Abel-transformed OH
emissions, (c) and (d) are mixture fraction fields (black=0; white=1) and
temperature fields (black=1000 K; yellow=3000 K). . . . . . . . . . . . . . 101
List of Figures 9

4.17 (a) Definition and (b) distribution of the integrated consumption rate of
oxygen O2 . (c) Normalized and cumulated consumption rate of oxygen
I O2 /I O2 ,max (defined in Eq. 4.13) . . . . . . . . . . . . . . . . . . . . . . 102
4.18 Axial evolution of (a) the momentum-flux ratio and (b) the velocity ratio. . 103
List of Tables

3.1 Species critical-point properties (temperature T , pressure P , molar volume


V and acentric factor ) and Schmidt numbers. . . . . . . . . . . . . . . . 36
3.2 Characteristic flow quantities . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Forward rate coecients in Arrhenius form k = AT n exp (E/RT ) for the
skeletal mechanism. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4.1 Characteristic flow quantities for the C60 operating point . . . . . . . . . . 84


Nomenclature iii

Nomenclature
Roman letters

Symbol Description Units Reference


Cp Specific heat capacity at constant pressure [J/(kg-K)]
Cv Specific heat capacity at constant volume [J/(kg-K)]
CS Constant of the standard Smagorinsky model [-]
Cw Constant of the Wale model [-]
Dk Molecular diusivity of the species k [m2 /s]
es Specific sensible energy [J/kg]
E Total energy [J/kg]
Ea Chemical activation energy [J]
gij Component (i, j) of the velocity gradient tensor [s1 ]
Jk,i Component (i) of the diusive flux vector of the [kg/(m2
species k s)]
J momentum-flux ratio (coaxial jets) [-]
k Turbulent kinetic energy [m2 /s2 ]
ndim Number of spatial dimensions [-]
P Pressure [N/m2 ]
qi Component (i) of the heat flux vector [J/(m2 s)]
R Universal gas constant [J/(mol
K)]
sij Component (i, j) of the velocity deformation ten- [s1 ]
sor
T Temperature [K]
u, ui Velocity vector/ component (i) [m/s]
Vic Component (i) of the correction diusion velocity [m/s]
Vik Component (i) of the diusion velocity of the [m/s]
species k
W Molecular weight [kg/mol]
Yk Mass fraction of the species k [-]
Z Mixture fraction [-]

Greek letters

Symbol Description Units Reference


ij Component (i, j) of the Kronecker delta [-]

Characteristic filter width [m]


h0f,k Formation enthalpy of the species k [J/kg]
k Kolmogorov length scale [m]
Thermal conductivity [J/(m-
K)]
iv Nomenclature

Dynamic viscosity [kg/(m-


s)]
Kinematic viscosity [m2 /s]
t Turbulent kinematic viscosity [m2 /s]
R Reduced pressure = P/Pc [-]
Equivalence ratio [-]
Density [kg/m3 ]
k Density of the species k [kg/m3 ]
R Reduced temperature = T /Tc [-]
c Characteristic chemical time scale [s]
conv Characteristic convective time scale [s]
ij Component (i, j) of the stress tensor [N/m2 ]
Artificial viscosity sensor [-]
kj Reaction rate of the species k in the reaction j [kg/(m3 s)]
T Heat release [J/(m3 s)]
ac Acentric factor used in real gas EOS []

Non-dimensional numbers

Symbol Description Reference


Da Damkhler
Le Lewis
Re Reynolds
Sc Schmidt

Subscripts

Symbol Description
0 Thermodynamic reference state
c Critical point property
k General index of the species
pb Pseudo-boiling point property

Superscripts

Symbol Description
f Filtered quantity
f Density-weighted filtered quantity

Abreviations
Nomenclature v

Acronym Description
DNS Direct Numerical Simulation
EOS Equation Of State
GH2 Gaseous Hydrogen
GOx Gaseous Oxygen
LES Large-Eddy Simulation
LOx Liquid Oxygen
LRE Liquid Rocket Engine
NSCBC Navier-Stokes Characteristic Boundary Condition
PR EOS Peng-Robinson Equation Of State
RANS Reynolds Averaged Navier-Stokes
RMS Root Mean Square
SRK EOS Soave-Redlich Kwong Equation Of State
Nomenclature vii

A Marina,
viii Nomenclature

Acknowledgments/Remerciement
En premier lieu, jaimerais remercier tous les membres du jury: le prsident Sbastien Can-
del et Patrick Chassaing pour avoir partag leur rudition en combustion et turbulence,
les rapporteurs Francis Dupoirieux et Julien Rveillon qui mont permis de prparer la
soutenance grce leurs remarques et questions trs dtailles, Didier Saucereau et Marie
Theron pour lintrt quils ont port mes travaux de thses et pour leurs clairages
sur les enjeux et problmatiques industrielles de R&T en propulsion liquide arospatiale.
Aussi, jaimerais remercier mon employeur, Snecma Vernon du groupe Safran et le CNES,
pour avoir financer mes travaux de recherche pendant trois ans (thse CIFRE). Je tiens
aussi remercier les centres de calculs franais et europens (CINES, IDRIS et PRACE),
pour les allocations dheures de calcul dont mes collaborateurs et moi ont pu bnficier
pour mener bien les simulations massivement parallles dont les rsultats sont prsentes
dans le prsent manuscrit.
Je tiens tout dabord remercier Thierry Poinsot et Bndicte Cuenot pour mavoir
accueilli au sein de lquipe combustion du CERFACS, quils dirigent (avec Laurent Gic-
quel et les autres sniors) avec talent. Je remercie aussi Thierry pour mavoir pouss
prsenter mes rsultats de thse dans des confrences internationales (ce qui ma permis
de rencontrer mon futur employeur !) et pour mavoir accueilli chez lui pendant quelques
jours agrables Portolla Valley. Je remercie aussi Marie Labadens, qui est un pilier
essentiel du bon fonctionnement de lquipe CFD.
Je remercie naturellement lquipe CSG (le trio Grard Dejean, Fabrice Fleury, Isabelle
DAst et lautre ct du couloir Patrick Laporte et Nicolas Monnier) pour leur support
inconditionnel et leurs rponses multiples mes demandes toujours plus farfelues. Je
remercie tout particulirement Sverine Toulouse et Nicole Boutet pour leur sympathie et
pour avoir partag ensemble la passion pour les animaux. Je noublie pas Chantal Nasri,
Michle Campassens et Lydia Otero pour leur amabilit et leur gentillesse naturelles.
Je tiens ensuite remercier tous les sages du CERFACS qui mont form au calcul
hautes performances et la mcanique des fluides numrique. En particulier Thomas
Schmitt, qui est en quelque sorte mon pre spirituel et qui a pris le temps de rpondre
toutes mes questions sur le supercritique et sur AVBPRG lors de mon stage de fin dtudes.
Je remercie aussi tout particulirement Gabriel Staelbach, qui ma normment guid et
aid dans la manipulation et la modification dAVBP, qui ma fait confiance et ma ouvert
les portes des centres de calculs nationaux et europens. Un grand merci Laurent Selle,
qui ma pris en main en deuxime anne en prononant la phrase : jaimerais quon
travaille ensemble. Travailler avec Laurent, qui est trs rigoureux et trs exigeant, ma
permis de me structurer et de mendurcir. Travailler avec Laurent a aussi t une grande
source de satisfaction, je me rappelle encore de son a dchire a !, en voyant mes premiers
rsultats de DNS. Je remercie aussi les autres sniors avec qui jai pu interagir : Olivier
Vermorel, Elonore Riber et Antoine Dauptain. Je remercie ensuite dautres anciens de
lquipe combustion du CERFACS, qui mont apport des clairages: Guilhem Lacaze,
Olivier Cabrit, Matthieu Leyko, Matthieu Boileau, Flix Jaegle, Marta Garcia, Nicolas
Lamarque, Simon Mendez et Florent Duchaine.
Aprs les sages, je remercie les apprentis sorciers, qui ont partag ces trois annes
Nomenclature ix

de thse, avec leurs bons et leurs mauvais moments. Je remercie tout dabord Benedetta
Franzelli, le guru ociel de Cantera, avant que Jean-Philippe Rocchi ne prenne le relais.
Merci vous deux et merci toi Jean-Phi, dtre rest jusqu la dernire minute, la
veille de ma soutenance, pour essayer dclaircir avec moi certains mystres des flammes
de diusion. Je remercie aussi Jorge Amaya, Camilo Silva et Pierre Wolf pour avoir jou
quelques notes de musique (entre autres) avec moi au dbut de la thse. Je remercie
Alexandre Neophytou, pour ses clairages sur les flammes triples et pour son humour so
british, Matthias Kraushaar pour son dvouement sans faille pour organiser des parties
de foot, Ignacio Duran pour les discussions en espagnol, Mario Falese pour sa curiosit
naturelle, Rmy Fransen et Stphane Jaur pour mavoir suivi dans laventure avbpedia,
Patricia Sierra et Victor Granet pour leur aide sur le dpart aux Etats-Unis et Alexandre
Eyssartier qui a d rsoudre la moiti des bugs sur les conditions limites cause de son
lien de parent avec genprofile !
Enfin, je remercie du fond du coeur mes collgues proches, devenu amis. Merci
Gregory Hannebique pour son soutien moral, pour les coups quil a essay de me porter
la boxe franaise, pour ses parcours top niveau (ou un peu plus) et pour son amiti,
du fat mec ! Merci Sebastian Hermeth pour ses bons conseils damis qui permettent
davancer dans la vie et pour ses cours dallemand, jetzt alles klar. Merci Georoy
Chaussonnet pour avoir russi supporter Greg et aussi pour avoir toujours le mot pour
rire. Merci Thomas Pedot, mon partenaire de bureau, qui ma permis davancer dans
la thse et a su me booster quand il fallait. Merci Ignacio Hernandez, alias Nacho (ou
p.....a), pour avoir emmen quotidiennement un parfum de banane dans notre bureau et
pour nous avoir fait dcouvrir la fte Madrid. Merci Jean-Franois Parmentier pour
mavoir redonn le got du fun dans la thse, avec des sujets de discussions totalement
dcousus et merci aussi de mavoir laiss finir ma thse en se limitant 15 minutes de
temps de paroles par jour.
Je remercie aussi mes parents, en particulier ma mre qui a dvou sa vie sa famille
et qui a pass tant dheures mes cts pour maider minstruire pendant mon enfance.
Je remercie aussi mon pre qui ma permis de dvelopper mon originalit (notamment
en rfrence au Loup et lagneau) et qui est source dinspiration pour moi. Cest aussi
grce mes parents formidables et leur ducation que jai pu russir un doctorat.
Je nai pas ralis ce travail de thse tout seul. Jai t port par ma femme, mon
amour, mon me soeur, Marina Beriat-Ruiz, qui ma soutenu contre vents et mares tout
au long de ces trois annes. Elle a su mencourager dans les moments o je ny croyais
plus, couter mes problmes de viscosit artificielle, comprendre ce qutait un maillage
et comprendre que ma passion me poussait passer mes soires et mes week-ends devant
un ordinateur au lieu de passer du temps avec celle que jaime. Le plus dur moments
ont t les derniers mois de rdaction, entremls avec lorganisation dun mariage, dun
dmnagement et dun dpart ltranger. Merci davoir supporter sur tes paules le
poids de mon absence.
Chapter 1
Introduction

Warning: the confidential parts of this thesis work have been removed and this manuscript
does not represent the entirety of the work done by Anthony Ruiz during his PhD.

1.1 Operating principle of Liquid Rocket Engines


The aim of a launcher is to safely send hu-
man beings or satellites into space. The eu-
ropean space launcher Ariane 5, depicted
in Fig. 1.1 is dedicated to the launch of
commercial satellites. The thrust that gen-
erates the lift-o and acceleration of the
launcher is produced by two dierent types Satellite(s)
of engine: solid-fuel rockets (or boosters)
and Liquid Rocket Engines (LREs), as
shown in Fig. 1.1. These two dierent
types of engine are complementary: the
boosters generate a very large thrust at Secondary stage
take-o but are depleted rapidly, whereas LRE
the main LRE generates a smaller thrust Oxidiser tank
but for a longer period of time.
Propulsion in boosters and LREs is ob- Boosters
tained through the reaction principle: the
ejection of fluid momentum from the noz-
zle of the engine, generates thrust. In this Fuel tank
system, the combustion chamber plays a
central role, since it converts chemical en-
ergy, stored in the reactants (also called
propellants) into kinetic energy.
Primary stage
Within the boosters, solid reactants are
LRE
arranged in a hollow cylindrical column, as
shown in Fig. 1.2. Combustion takes place
at the inner surface of the cylindrical col- Figure 1.1: The main components of the Ar-
umn, whose volume decreases due to abra- iane 5 european space launcher.
sion until total depletion of reactants. The
operating principle of a booster is fairly
simple, however, complex and coupled phenomena take place within it. For instance,
2 Chapter 1. Introduction

the failure of the space shuttle challenger, which lead to the deaths of its seven crew
members in 1986, was caused by the leakage of hot gases through the rocket casing, due
to abnormal ignition stress. These hot gases caused a structural failure of the adjacent
tank [William P. Rogers 1986]. In order to understand these complex phenomena, fully-
coupled simulation of reacting fluid flow and structure oscillations are now possible. For
instance, see [Richard & Nicoud 2011], among others.

Figure 1.2: Components of a booster.

The operating principle and the type of combustion are totally dierent in a LRE.
Fuel and oxidizer are stored in a liquid state, in separate tanks, as shown in Fig. 1.1.
Turbopumps inject these non-premixed reactants into the combustion chamber, where
they mix through molecular diusion and turbulence, and burn, which expands the gas
mixture.
Figure 1.3 shows the operating principle of Vulcain 2 [Snecma 2011]. The torque
required to drive the turbopumps is generated by the flow of combustion products from
the gas generator, which is a small combustion chamber, through turbine blades. A small
amount of this torque is used to inject reactants into the gas generator itself. Finally,
to start the gas generator, a solid combustion starter drives the turbopumps for a few
seconds.
In general, the combustion chamber of a LRE is composed of a large number of coaxial
injectors, arranged on a single plate as shown in Fig. 1.4. This compact arrangement of
multiple flames enables a large delivery of thermal power in a small volume. The thermal
power delivered by Vulcain 2 is 2.5 GW and the volume of the combustion chamber is
20 L. This yields a very large power density of 50 GW/m3 , which is equivalent to the
power of 50 nuclear units in 1 cubic meter.
1.1. Operating principle of Liquid Rocket Engines 3

LH2 reservoir LOX reservoir

Gaz generator

> 500 coaxial injectors

Turbopumps
Combustion Chamber

Nozzle

Liquid Oxygen (LOX)


Liquid Hydrogen (LH2)
Combustion products

Figure 1.3: Operating principle of the Vulcain 2 engine [Snecma 2011].

coaxial
injectors

igniter
Figure 1.4: Injection plate of the Vulcain 2 LRE, composed of 566 coaxial injectors [As-
trium 2011]
4 Chapter 1. Introduction

1.2 Combustion in LREs


This thesis work is focused on flame dynamics inside the combustion chambers of LREs
(gas generator and main combustion chamber). The detailed understanding of combustion
physics inside these critical propulsive organs can provide useful guidelines for improving
the design of LREs.
This section reflects the state of the art in the field of combustion, at operating con-
ditions relevant to LREs. There are several reviews on the topic, which are synthesized
and complemented herein. The reader is referred to the following publications for more
details on:

experimental results [Candel et al. 1998,Haidn & Habiballah 2003,Candel et al. 2006,
Oschwald et al. 2006, Habiballah et al. 2006]

theory and modeling of turbulent supercritical mixing [Bellan 2000,Bellan 2006] and
combustion [Oefelein 2006]

1.2.1 Preliminary Definitions


1.2.1.1 Supercritical and transcritical injection
In the reservoirs of a launcher, reactants are stored at a very low temperature, typically
100 K for oxygen and 20 K for hydrogen. This gives the reactants a very large density,
which allows to reduce the dimension of the reservoirs and increases the possible payload.
In the combustion chamber, heat release expands the gas mixture and increases its tem-
perature. For instance, the adiabatic flame temperature of a pure H2 /O2 flame at 10 MPa
is Tadiab = 3800 K.
Since the main LRE of a launcher is started at the ground level, the initial pressure
inside the combustion chamber is around 0.1 MPa. After ignition, the pressure rises until
the nominal value is reached, which is approximately equal to 10 MPa.
Thus, the path followed by an O2 fluid parcel from the reservoir to the flame zone in
a LRE at nominal operating conditions is schematically represented by the arrow on the
top Fig. 1.5, which shows the states of matter, as well as isocontours of heat capacity,
in a T-P diagram for pure oxygen. Below the critical pressure (Pc,O2 = 5.04 MPa), the
boiling line (or saturation curve) is the frontier between the liquid and gaseous state.
By extension, at supercritical pressure, the pseudo-boiling line (Tpb ,Ppb ) follows the local
maxima of heat capacity and is thus the solution of the following system:

P > Pc

Cp
= 0 (1.1)
T P

The injection of oxygen inside a LRE is called transcritical because of the crossing of
the pseudo-boiling temperature, which is accompanied with a change of thermophysical
quantities (such as density or diusion coecients) that vary from liquid-like (for T < Tpb )
to gas-like (for T > Tpb ) values.
1.2. Combustion in LREs 5

Transcritical combustion
100
Pseudo Boiling Line
80 Supercritical Fluid
60
P [bar]

40

Liquid Gas
20 Boiling Line

100 150 200 250 300


T [K]

Figure 1.5: Heat capacity isocontours computed with the Soave-Redlich-Kwong equation
of state (see Eq. 2.44): white = 103 J/K/kg; black=104 J/K/kg.

1.2.1.2 Scalar dissipation rate and Damkhler number


For non-premixed combustion, there is a balance between the mass flux of reactants to
the flame front and the burning rates of these reactants. The mass flux of reactants
to the flame front is directly related to species gradients, which determine the intensity
of molecular fluxes. The scalar dissipation rate is a measure of the molecular diusion
intensity and defines the inverse of a fluid time scale:
= 2D|Z|2 (1.2)
where D is the thermal diusivity (m2 /s), and Z is the mixture fraction [Peters 2001].
The mixture fraction is equal to the mass fraction of the H atom for pure H2 /O2
reacting flows [Poinsot & Veynante 2005] and depends on the number of species contained
in the gas mixture. For instance, for the detailed chemical scheme used in Chap. 3, the
mixture fraction reads:

YH 2 YH2 O YH YOH YH2 O2 YHO2
Z = WH 2 +2 + + +2 + (1.3)
WH2 WH2 O WH WOH WH2 O2 WHO2
where Yk and Wk are the mass fraction and the molecular weight of the species k.
The ratio of the flow time tf low and the chemical time tchem is the Damkhler number:
tf low
Da = (1.4)
tchem
A
= (1.5)

6 Chapter 1. Introduction

where A is a typical Arrhenius pre-exponential of the chemical kinetic scheme.

1.2.1.3 Mixture ratio


The mixture ratio E is the ratio of the injected mass flow rate of oxidizer mO and fuel
mF inside a combustion chamber:
mO
E= (1.6)
mF
In a LRE, fuel is generally injected in excess to protect the chamber walls from oxidation,
consequently, the mixture ratio is generally smaller than the stoichiometric mass ratio s:
O WO
s= (1.7)
F WF

where O and F are the stoichiometric coecients of an overall unique reaction [Poinsot
& Veynante 2005]:

F F + O O Products (1.8)

For H2 /O2 combustion, s = 8 and for CH4 /O2 combustion, s = 4.

1.2.2 Experimental studies


Flame structure at subcritical pressure
Experimental studies of flame patterns in rocket conditions were first conducted at
subcritical pressure, with coaxial injectors feeded with H2 /LOx. The pioneering experi-
mental works of [Herding et al. 1996, Snyder et al. 1997, Herding et al. 1998] have allowed
to determine some key characteristics of cryogenic propellant combustion that are now
well established, using advanced optical diagnostics (OH together with Planar Laser In-
duced Fluorescence (PLIF)). One of the most important feature that has been discovered
is the anchoring of the flame on the injector rim, independently of operating conditions
for H2 /O2 .
The OH technique use the physical principle that the flame emits light that is com-
posed of wavelengths representative of the local gas composition. By sensing the light at
a given wavelength, the location of species involved in chemical kinetics can be revealed
within the flow field. The excited hydroxyl radical OH is a suitable flame indicator since
it is an intermediate species and is thus only present at the flame front, whereas H2 O is
also present in the burnt gases. The OH signal is an integration along the line-of-sight of
the chemiluminiscence and although it provides an overview of the instantaneous reacting
zone inside the flow field, it does not provide a precise view of the flame shape, which is
an important information for the design of engines and for validation purposes.
In the PLIF technique, a laser sheet excites an intermediate species (OH for instance)
that emits light to relax towards an unexcited state. This technique was used at subcritical
pressure in [Snyder et al. 1997] to observe instantaneous flame fronts and study the flame
shape and stabilization in reacting LOx/H2 flows. Because the signal-to-noise ratio is
1.2. Combustion in LREs 7

deteriorated with increasing pressure, the PLIF technique was only used at moderate
pressures (less than 1 MPa) in the 1990s. Thus, the only technique available for the study
of high-pressure reacting flow at that time was OH emission.
A major breakthrough in the experimental study of cryogenic propellant combustion
was made in [Herding et al. 1998], where the Abel transform was first use to retrieve
a cut through a time-averaged OH signal, with the assumption of axi-symmetry. This
technique was then used in experimental studies of high-pressure reacting flows, to pre-
cisely monitor the eects of operating conditions and geometry of injectors on the flame
pattern [Juniper et al. 2000, Singla et al. 2005].

Subcritical and supercritical characteristics of atomization


Since nitrogen is an inert species and bears similar characteristics to oxygen (mo-
lar mass and critical point), it has been used as an alternative species for the study
of the characteristics of atomization of jets inside LREs, at subcritical and supercriti-
cal conditions. Most non-reacting experimental results have been obtained at the Air
Force Research Laboratory (AFLR) and the German Aerospace Center (DLR) [Chehroudi
et al. 2002a, Chehroudi & Talley 2001, Chehroudi et al. 2002b, Mayer et al. 1998b, Mayer
& Branan 2004, Branam & Mayer 2003, Mayer & Smith 2004, Mayer et al. 1996, Mayer
et al. 2000,Mayer et al. 2001,Mayer et al. 1998a,Mayer & Tamura 1996,Mayer et al. 2003].
It has been shown that the crossing of either the critical temperature or the critical pres-
sure impacts atomization. At subcritical conditions, because of the joint action of aerody-
namic forces and surface tension, ligaments and droplets are formed at the surface of liquid
jets. At supercritical conditions (either T > Tc and/or P > Pc ), surface tension vanishes
and droplets are absent from the surface of jets. Instead, in the case of a transcritical
injection (see Fig. 1.5), a diusive interface between dense and light fluid develops, where
waves or comb-like structures can form. Figure 1.6 shows a time sequence of a trans-
critical N2 /supercritical He (inert substitute for hydrogen) mixing layer, where traveling
waves are observed in the diuse interface [Teshome et al. 2011]. The luminance of a
shadowgraph is sensitive to gradients of refractive index, which allow to localize interfaces
between jets having dierent densities.
The normalization of the injection temperature and the chamber pressure by the crit-
ical coordinates defines the reduced temperature R and the reduced pressure R
T
R = (1.9)
Tc
P
R = , (1.10)
Pc
and allows to identify the atomization regime.
It is thus expected that during the ignition transient of LRE, the mixing between
oxygen and hydrogen is greatly impacted by the crossing of the critical pressure.

Flame structure at supercritical pressure


The first experimental results obtained at supercritical pressure, at operating conditions
that are relevant to LREs, have been obtained in [Juniper et al. 2000]. The same advanced
8 Chapter 1. Introduction

GHe

LN2

Figure 1.6: Shadowgraphs at successive instants (time between frames is 0.1 ms) of a
transcritical N2 /supercritical He mixing layer [Teshome et al. 2011].

optical diagnostics (OH and PLIF) that were developed and used at subcritical pressure
have successfully been applied to high-pressure combustion. An instantaneous OH image
is shown in Fig. 1.7(a), a time-averaged OH image and its Abel transform is shown in
the top and the bottom of Fig. 1.7(b), respectively. One can clearly see the progress that
has been made in optical diagnostics using the Abel transform, where highly reacting
zone can be observed at the inner shear layer, in the near-injector region, and where an
opening turbulent flame brush can be observed further downstream.
Other important features of cryogenic propellant combustion were discovered in later
work [Juniper & Candel 2003a], where a stabilization criterion has been devised, which
essentially states that the lip height should be larger than the flame thickness to enable
flame anchoring on the injector rim and overall flame stabilization.
In [Singla et al. 2005], methane was used instead of hydrogen, which have enabled
the identification of a new flame structure under doubly transcritical injection conditions:
two separate flames were observed at the inner and the outer shear layers of a coaxial
1.2. Combustion in LREs 9

Time-averaged

Abel-transformed

(a) (b)

Figure 1.7: Flame shape visualizations using OH , in a coaxial LOx/GH2 injector [Ju-
niper 2001]: (a) instantaneous image, (b) top: time-averaged image; bottom: Abel-
transformed image. The pressure is 7 MPa.

injection.
In comparison with the Abel transform of a time-averaged OH signal, the PLIF pro-
vides an instantaneous visualization of the flame front and enables the study of unsteady
features of combustion, such as moving flame tip at the stabilization point or flame front
wrinkling due to turbulent motions. However, [Singla et al. 2006,Singla et al. 2007] are the
only experimental studies that have successfully used the PLIF technique at supercritical
pressures, in transcritical conditions, as shown in Fig. 1.8. On the latter figure, there is
no signal from the upper flame sheet as most of the light from the laser is absorbed by
the OH distribution on the lower side facing the sheet transmission window.
A close-up view of the near-injector region of PLIF images have shown that transcrit-
ical LOx/GH2 flames are stabilized on the injector rim, which indicates that in this type
of flames, the Damkhler number (see Eq. 1.4) is suciently large to prevent extinction.

More recently, other experimental studies of reacting fluid flows in conditions relevant
to LRE, have been conducted in [Locke et al. 2010], at an unprecedented data acquisition
rate, using shadowgraphy. Since the flame zone is a region with large density gradients
(because of thermal expansion), it can be visualized using shadowgraphy. This has allowed
to observe new features of cryogenic propellant combustion, such as the emission of oxygen
pockets from the inner jet, as shown in Fig. 1.9, which might play an important role on
unsteady heat release and possibly combustion instabilities.
10 Chapter 1. Introduction

Figure 1.8: OH PLIF image of a LOx/GH2 cryogenic flame [Singla et al. 2007].

Figure 1.9: Shadowgraphs of a transcritical H2 /O2 reacting flow at 6 MPa, taken at


successive instants (time between frames is 0.25 ms) [Locke et al. 2010].
1.2. Combustion in LREs 11

Influence of operating conditions and geometry of coaxial injector


Figure 1.10, extracted from [Schumaker & Driscoll 2009], shows a schematic of the
near-field mixing layers in a coaxial jet. An inner shear layer is located at the interface

he
h
ri

Figure 1.10: Schematic of coaxial jet injector and the near-field mixing layers [Schumaker
& Driscoll 2009]

between the high-speed outer light jet and the low-speed inner dense jet. The outer shear
layer lies in between the outer hydrogen and the recirculated gases of the combustion
chamber.
An ecient coaxial injector should maximize turbulent mixing of reactants so that the
resulting flame length and temperature stratification of burnt gases is as small as possible.
The Reynolds number compares a convection time c = L/u and a diusion time
d = L2 /:
d
Re =
c
LU
= (1.11)

with the density (kg/m3 ), L a characteristic length scale (m), U a velocity (m/s) and
the dynamic viscosity (Pa.s). For a coaxial injector, a Reynolds number is defined for
each of the inner and outer streams:
O 2 d i Ui
Rei = (1.12)
O 2
H2 he Ue
Reo = (1.13)
H 2
12 Chapter 1. Introduction

where the reference lengths are the inner diameter di (= 2 ri ) and the outer channel
height he . These Reynolds numbers characterize the fluid flow inside the feeding lines,
immediately upstream of the injection plane. Because of the high injection velocities inside
a LRE, the Reynolds number is large, typically above 105 , which ensures the jets issuing
from the coaxial injector rapidly transition towards turbulence. Thus, in the context of
LREs, the Reynolds number is generally not an influential parameter.
Experimental studies [Snyder et al. 1997, Lasheras et al. 1998, Favre-Marinet & Ca-
mano Schettini 2001] have identified the density ratio R and the momentum-flux ratio J
as the two numbers influencing mixing eciency in coaxial configurations:
O2
R = (1.14)
H 2
H2 UH2 2
J = . (1.15)
O2 UO2 2

The velocity ratio appears to be less influential:


uH 2
Ru = , (1.16)
u O2
(1.17)

and can be deduced from the density and momentum-flux ratio: Ru = J R . [Favre-
Marinet & Camano Schettini 2001] have determined that the potential core length is
proportional to J 1/2 for variable-density non-reacting jets, over a very wide range of
values (0.1 < J < 50), spanning typical values in LRE (1-20).
The geometry of the coaxial injector also influences combustion eciency. In partic-
ular, it has been shown in [Juniper et al. 2000, Juniper & Candel 2003b] that recessing
the inner LOx tube from the outer H2 tube increases combustion eciency. This point is
treated in more details in Chap. 4.
1.2. Combustion in LREs 13

1.2.3 Numerical studies


Temporal mixing layers
The pioneering work of [Bellan 2000, Okongo & Bellan 2002b, Bellan 2006] relied on
Direct Numerical Simulations (DNSs) of temporal non-reacting mixing layers to study
the impact of real-gas thermodynamics on the characteristics of turbulent mixing. These
studies also constituted databases that were used a priori and a posteriori to determine
the validity of closure terms for Large Eddy Simulation (LES). These studies lead to the
following major findings:

shear-induced instability at the interface between a dense and a light fluid give
rise to larger velocity fluctuations in the light fluid than in the dense fluid and the
density gradient redistributes turbulent kinetic energy from the perpendicular to
the parallel direction of the interface [Okongo & Bellan 2002b]

because real-gas equation of states are non-linear, an additional subgrid-scale term,


related to the filtering of pressure, requires closure in the LES framework [Bel-
lan 2006].

Recently, [Foster 2009, Foster & Miller 2011] have studied a temporal reacting mixing
layer between H2 and O2 , at a large Reynolds number, and extended the a priori analysis
of subgrid-scale turbulent fluxes to reacting conditions. However, these studies have not
yet determined an adequate closure of subgrid-scale terms that could be used in LES
simulations of realistic configurations.

Round jets
Several numerical studies attempted to predict the dierence in turbulent mixing in-
duced by the crossing of the critical pressure or the critical temperature.
In [Zong et al. 2004,Zong & Yang 2006], the eect of chamber pressure on supercritical
turbulent mixing was investigated with LESs of the round jet experiments conducted with
N2 in [Chehroudi et al. 2002a]. An increase in the ambient pressure results was shown
to provoke an earlier transition of the jet into the self-similar regime, because of reduced
density stratification. Figure 1.11 shows radial distributions of normalized density at
various axial locations, showing the existence of a self-similar region in supercritical jets.
In [Schmitt et al. 2010b], the eect of injection temperature on supercritical turbulent mix-
ing was investigated with LESs of the round jet experiments conducted with N2 in [Mayer
et al. 2003]. The comparison between the mean axial density profiles obtained in the
numerical simulation and the experimental measurements showed good agreement. The
increase of the injection temperature above the pseudo-boiling temperature (see Eq. 1.1),
lead to an early transition of the jet into the self-similar regime.
RANS simulations of the cryogenic round jet experiments were also conducted in
[Mayer et al. 2003, Kim et al. 2010], and were able to qualitatively reproduce the mean
density profiles from the experiment.
14 Chapter 1. Introduction

Figure 1.11: Radial distributions of normalized density at dierent axial locations (T =


300 K, uinj = 15 m/s, Tinj = 120 K, Dinj = 254 m) [Zong & Yang 2006]
1.2. Combustion in LREs 15

Flame stabilization region


To the authors knowledge, [Oefelein & Yang 1998] is the first numerical work that
focused on the stabilization point of a cryogenic LOx/GH2 flame. In these numerical
simulations, the flames were stabilized at the injector rim for supercritical and trans-
critical mixing, as shown in Fig. 1.12. The stabilizing eect of the transcritical density
gradient was evidenced from these simulations, as visually observed when comparing the
transcritical and the supercritical flame shapes in Figs. 1.12(a) and 1.12(b), respectively.
The stabilization of the flame tip at the injector rim has later been confirmed in other
experimental [Juniper et al. 2000, Singla et al. 2007] and numerical studies [Juniper &
Candel 2003a].

(a) (b)

Figure 1.12: Contours of temperature for the near-field region for (a) supercritical and
(b) transcritical mixing [Oefelein & Yang 1998].

In [Zong & Yang 2007] the study of the near-field region of a coaxial injector fed-in
with transcritical LOx/GCH4 , in the same configuration as in [Singla et al. 2007], also
showed a flame stabilized at the injector rim.

Transcritical flame structure


Laminar counterflow diusion flames have been studied in the literature, to scrutinize
the transcritical chemical structure of LOx/GH2 and LOx/GCH4 flames.
In [Juniper et al. 2003], the study of a diusion flame developing between gaseous
hydrogen and liquid oxygen at atmospheric pressure, showed that the flame structure
bears similar characteristics as a flame located in between gaseous reactants. It was also
shown that the extinction strain rate is large for LOx/GH2 (larger than 5 105 s1 ), and
increases with pressure.
In [Pons et al. 2008] and [Ribert et al. 2008], it was shown for LOx/GH2 and
LOx/GCH4 that the heat release rate per unit flame surface increases with the square
root of strain rate and pressure, while the extinction strain rate evolve quasi-linearly with
pressure. These studies also showed that a transcritical flame structure is very similar to
a supercritical flame structure.
16 Chapter 1. Introduction

In the previous studies, the chemical reaction rates were assumed to bear a standard
Arrhenius form. In [Giovangigli et al. 2011], real-gas thermodynamics derivations were
used to compute chemical production rates from chemical potentials. It was found that
the non-ideal chemical production rates may have an important influence at very high
pressure (typically above 20 MPa).

Coaxial jet flames


Only a few numerical studies have used LES as a tool for detailed understanding of
flame dynamics inside LREs. For instance, [Masquelet et al. 2009] led a 2D-axisymmetric
simulation of a multiple-injector combustor, and focused on wall heat flux induced by
combustion. Transcritical LESs of coaxial injectors were conducted in [Schmitt et al. 2009,
Schmitt et al. 2010a] and [Matsuyama et al. 2006,Matsuyama et al. 2010], where the flow
complexity was captured with great precision and good agreement with experimental
data, as shown in Fig. 1.13.
1.2. Combustion in LREs 17

(a)

(b)

Figure 1.13: LES computations of transcritical jet flames: (a) T = 1000 K iso-surface
colored by axial velocity in a reacting transcritical LOx/GH2 flow [Matsuyama et al. 2010].
(b) Visualization of a transcritical LOx/GCH4 flame: (top) direct visualization from
experiment [Singla 2005]; (bottom) T = (T max + T min )/2 isosurface from LES [Schmitt
et al. 2010a].
18 Chapter 1. Introduction

1.3 Study Plan


The objective of the present thesis work is to study the characteristics of reacting flows at
operating conditions and in configurations relevant to LREs, with high-fidelity unsteady
numerical simulations. Figure 1.14 shows a schematic view of the combustion chamber of
a LRE, and the dierent study levels considered herein.

1 - Near-injector region
1
2 - Single injector

3 - Multiple injectors
2
3

Figure 1.14: Schematic of a LRE combustion chamber, showing the dierent levels of
study considered in the present thesis work.

1 - Near-injector region
In Chap. 3, the stabilization region of a transcritical LOx/GH2 flame is simulated with
Direct Numerical Simulation with an unprecedented spatial resolution. The non-reacting
and reacting flow characteristics are compared, and information is provided on the flame
stabilization mechanism and on the turbulent flame structure in LREs.

2 - Single injector
In Chap. 4, a transcritical LOx/GH2 jet flame stabilized at the rim of a coaxial injec-
tor, is simulated with LES. The numerical results provide a detailed view of the turbulent
mixing process in such a configuration. Then, an important design parameter (the in-
ner recess length of the oxygen tube) is varied and the eects of this parameter on the
flame dynamics is investigated. Numerical results are compared with experimental and
theoretical results for validation.

3 - Multiple Injectors
A preliminary LES result of a multiple-injector combustor is shown in Chap. 5. The final
aim of such configurations is to understand how multiple flames interact under smooth
operating conditions or under transverse acoustic waves, as often encountered prior to
combustion instability in LREs.
Chapter 2
Governing Equations, Thermodynamics
and Numerics

Contents
2.1 Navier-Stokes Equations . . . . . . . . . . . . . . . . . . . . . . . . 20
2.1.1 Species diusion flux . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.2 Viscous stress tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.3 Heat flux vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.1.4 Transport coecients . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.2 Filtered Equations for LES . . . . . . . . . . . . . . . . . . . . . . . 25
2.2.1 The filtered viscous terms . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.2 Subgrid-scale turbulent terms for LES . . . . . . . . . . . . . . . . . 26
2.3 Models for the subgrid-stress tensor . . . . . . . . . . . . . . . . . 28
2.3.1 Smagorinsky model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.3.2 WALE model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4 Real-Gas Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . 28
2.4.1 Generalized Cubic Equation of State . . . . . . . . . . . . . . . . . . 29
2.4.2 Primitive to conservative variables . . . . . . . . . . . . . . . . . . . 29
2.5 CPU cost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

This chapter presents the conservation equations implemented in the AVBP numerical
solver (used in later DNS and LES studies (Sec. 3 to Sec. 5)).
The AVBP solver is co-developed by CERFACS and IFP Energies Nouvelles. It uses
unstructured meshes and finite-volume and finite-element high-order schemes to discretize
the direct and filtered Navier-Stokes equations in complex geometries for combustion and
aerodynamics.
AVBP has been used in many dierent applications, ranging from gas turbines [Selle
et al. 2004, Roux et al. 2005, Wolf et al. 2009] and piston engines [Vermorel et al. 2009,
Enaux et al. 2011], to scramjets [Roux et al. 2010]. An overview of the solver properties
and some recent industrial applications of AVBP is presented in [Gourdain et al. 2009a,
Gourdain et al. 2009b].
The description of numerical schemes in the unstructured framework is omitted and
may be found in the handbook of the AVBP solver [AVBP 2011] or in previous thesis
20 Chapter 2. Governing Equations, Thermodynamics and Numerics

manuscripts from CERFACS. See [Lamarque 2007, Jaegle 2009, Senoner 2010], among
others.

2.1 Navier-Stokes Equations


Throughout this document, it is assumed that the conservation equations for a super-
critical fluid flow are similar to a perfect-gas variable-density fluid. The only dierence
between the two is that they are coupled to dierent equations of state. This single-phase
formulation assumes that surface tension is negligible in front of aerodynamic forces (infi-
nite Weber number), which appears to be an overall-correct approximation in the light of
experimental studies of cryogenic flows (see Sec. 1.2.2). However, the fundamental studies
of [Giovangigli et al. 2011] have shown that even at supercritical pressures, a fluid mixture
can be unstable and phase change can still occur in certain cases. There are thus various
levels of accuracy in the description of supercritical flows and in a simplified approach,
it is assumed here that the supercritical fluid mixture is always stable so that no phase
change occur.
Throughout this part, the index notation (Einsteins rule of summation) is adopted for
the description of the governing equations. Note however that index k is reserved to refer
to the k th species and will not follow the summation rule unless specifically mentioned or

implied by the sign.
The set of conservation equations describing the evolution of a compressible flow with
chemical reactions reads:
u i
+ ( ui uj ) = [P ij ij ] (2.1)
t xj xj

E
+ ( E uj ) = [ui (P ij ij ) + qj ] + T (2.2)
t xj xj

k
+ (k uj ) = [Jj,k ] + k . (2.3)
t xj xj
In equations 2.1 to 2.3, which respectively correspond to the conservation laws for
momentum, total energy and species, the following symbols (, ui , E, k ) denote the
density, the velocity vector, the total energy per unit mass (E = ec + e, with e the sensible
energy) and the density of the chemical species k: k = Yk for k = 1 to N (where N is
the total number of species); Yk being the mass fraction of the k th species. P denotes the
pressure (see Eq. 2.4.1), ij the stress tensor (see Eq. 2.12), qj the heat flux vector (see
Eq. 2.14) and Jj,k the vector of the diusive flux of species k (see Eq. 2.11). The source
term in the species transport equations (k in Eq. 2.3) comes from the consumption or
production of species by chemical reations. The source term in the total energy equation
(T in Eq. 2.2) is the heat release rate, which comes from sensible enthalpy variation
associated to species source terms:
N

T = (k h0f,k ) (2.4)
k=1
2.1. Navier-Stokes Equations 21

with h0f,k , the mass enthalpy of formation of species k.


The fluxes of the Navier-Stokes equations can be split into two parts:
Inviscid fluxes:

ui uj + P ij
(E + P ij ) uj (2.5)
k u j
Viscous fluxes:

ij
(ui ij ) + qj (2.6)
Jj,k

2.1.1 Species diusion flux


In multi-species flows the total mass conservation implies that:
N

Yk Vik = 0 (2.7)
k=1

where Vik are the components in directions (i=1,2,3) of the diusion velocity of species k.
They are often expressed as a function of the species molar concentration gradients using
the Hirschfelder-Curtis approximation:
Xk
Xk Vik = Dk , (2.8)
xi
where Xk is the molar fraction of species k: Xk = Yk W/Wk . In terms of mass fraction,
the approximation 2.8 may be expressed as:

Wk Xk
Yk Vik = Dk (2.9)
W xi
Summing Eq. 2.9 over all ks shows that the approximation 2.9 does not necessarily
comply with equation 2.7 that expresses mass conservation. In order to achieve this, a
correction diusion velocity V c is added to the diusion velocity Vik to ensure global mass
conservation [Poinsot & Veynante 2005]:
N
Wk Xk
Vic = Dk (2.10)
k=1
W xi

and computing the diusive species flux for each species k as:

Wk Xk
Ji,k = Dk Yk V ic
(2.11)
W xi
Here, Dk are the diusion coecients for each species k in the mixture (see section
2.1.4). Using equation Eq. 2.11 to determine the diusive species flux implicitly verifies
Eq. 2.7.
22 Chapter 2. Governing Equations, Thermodynamics and Numerics

2.1.2 Viscous stress tensor


The stress tensor ij is given by:

1
ij = 2 Sij ij Sll (2.12)
3

where Sij is the rate of strain tensor and is the dynamic viscosity (see section 2.1.4).

1 ui uj
Sij = + (2.13)
2 xj xi

2.1.3 Heat flux vector


For multi-species flows, an additional heat flux term appears in the diusive heat flux.
This term is due to heat transport by species diusion. The total heat flux vector then
takes the form:
N
N
T Wk Xk T
qi = Dk c
Yk Vi hk = + Ji,k hk(2.14)
x W x x
i k=1
i

i
k=1
Heat conduction Heat flux through species diusion

where is the heat conduction coecient of the mixture (see section 2.1.4) and hk the
partial-mass enthalpy of the species k [Meng & Yang 2003].
Dufour terms in Eq. 2.14 and Soret terms in Eq. 2.11 have been neglected, because
they are thought to be second-order terms for the type of flow investigated here (in a
DNS of reacting H2 /O2 , [Oefelein 2006] showed they were negligible in front of the other
terms), which would add an unnecessary level of complexity.

2.1.4 Transport coecients


In CFD codes for perfect-gas multi-species flows the molecular viscosity is often assumed
to be independent of the gas composition and close to that of air. In that case, a simple
power law can approximate the temperature dependency of a gas mixture viscosity.
b
T
= c1 (2.15)
Tref

with b typically ranging between 0.5 and 1.0. For example b = 0.76 for air.
For transcritical combustion, it is not possible to describe the liquid-like viscosity
and the gas-like viscosity with the same expression. Instead, the Chung model is used to
compute dynamic viscosity as a function of T and k [Chung et al. 1984,Chung et al. 1988].
The dynamic viscosity decreases with temperature in a liquid and increases with tem-
perature in a gas. This is illustrated by the evolution of dynamic viscosity and thermal
conductivity with temperature at 10 MPa for pure oxygen, as shown in Fig. 2.1. In this
figure, the transport coecients for heat and momentum are computed with the method
2.1. Navier-Stokes Equations 23

4
2 x 10
SRK
[Pa.s] NIST

0
0 500 1000 1500 2000 2500 3000
T [K]
(a)

0.2

0.15
[W/m/K]

SRK
0.1
NIST

0.05

0
0 500 1000 1500 2000 2500 3000
T [K]
(b)

Figure 2.1: Transport coecients for O2 at 100 bar, showing the liquid-like to gas-like
transition of thermo-physical properties . a) Dynamic viscosity b) Thermal conductivity.

presented in [Chung et al. 1984,Chung et al. 1988] which compares favorably to the NIST
database [Lemmon et al. 2009].
The computation of the species diusion coecients Dk is a specific issue. These
coecients should be expressed as a function of the binary coecients Dij obtained from
kinetic theory [Hirschfelder et al. 1954]. The mixture diusion coecient for species k,
Dk , is computed as [Bird et al. 1960]:

1 Yk
Dk = N (2.16)
j=k Xj /Djk

The Dij are functions of collision integrals and thermodynamic variables. In addition
to the fact that computing the full diusion matrix appears to be prohibitively expensive
in a multidimensional unsteady CFD computation, there is a lack of experimental data to
validate these diusion coecients. Therefore, a simplified approximation is used for Dk .
The Schmidt numbers Sc,k of the species are supposed to be constant so that the binary
diusion coecient for each species is computed as:

Dk = (2.17)
Sc,k
24 Chapter 2. Governing Equations, Thermodynamics and Numerics

This is a strong simplification of the diusion coecients since it is well-known that the
Schmidt numbers are not constant in a transcritical flame and can vary within several
orders of magnitude between a liquid-like and a gas-like fluid [Oefelein 2006]. Models exist
to qualitatively take into account this variation [Hirschfelder et al. 1954, Bird et al. 1960,
Takahashi 1974], although no experimental data is available to validate them. However,
the eects of the constant-Schmidt simplification on the flame structure is actually very
small, as shown later on in Fig. 3.6, and thus appears to be reasonable for the present
studies of combustion.
The diusion coecients of heat (Dth = /Cp ) and momentum ( = /) are sub-
mitted to large variations within the flow field, which have to be taken into account.
Due to the peak of Cp at the pseudo-boiling point, heat diusivity is lower than species
diusivity, as shown in Fig. 2.2. The Lewis numbers (Lek = Dth /Dk ) are below one, and
species gradients are likely to be smaller than temperature gradients in a transcritical
mixing layer.

2 Le
C /C
p p,0
1.5

0.5

0
0 100 200 300 400 500 600 700 800 900 1000
T [K]

Figure 2.2: Evolution of the Lewis number and normalized heat capacity with temper-
ature. Cp,0 = 1700 J/K/kg. The peak of Cp at the pseudo-boiling point creates a local
minimum in the Lewis number.
2.2. Filtered Equations for LES 25

2.2 Filtered Equations for LES


The idea of Large-Eddy Simulation (LES) is to solve the filtered Navier-Stokes equation,
modeling the small scales of turbulence, which are assumed to have a mere dissipation
role of the turbulent kinetic energy. The derivation of the LES equations starts with the
introduction of filtered quantities.
The filtered quantity f is resolved in the numerical simulation whereas the subgrid-
scales f = f f are modeled. For variable density , a mass-weighted Favre filtering of
f is introduced such as:

f = f (2.18)

The conservation equations for LES are obtained by filtering the instantaneous equa-
tions 2.1, 2.2 and 2.3:
ui
+ ( ui uj ) = [P ij ij ij t ] (2.19)
t xj xj


E
+ uj ) = [ui (P ij ij ) + qj + qj t ] + T
( E (2.20)
t xj xj

Yk t
+ ( Yk uj ) = [Jj,k + Jj,k ] + k (2.21)
t xj xj
In equations 2.19, 2.20 and 2.21, there are now four types of terms to be distinguished:
inviscid fluxes, viscous fluxes, source terms and subgrid-scale terms.
In this section, the subgrid-scale models for heat, mass and momentum fluxes are
assumed to bear the same form as in the perfect-gas case.
Inviscid fluxes:
These terms are equivalent to the unfiltered equations except that they now contain
filtered quantities:


ui uj + P ij
E uj + P uj ij (2.22)
k uj
Viscous fluxes:
The viscous terms take the form:


ij
(ui ij ) + qj (2.23)
Jj,k
Filtering the balance equations leads to unclosed quantities, which need to be modeled,
as presented in Sec. 2.2.2.
26 Chapter 2. Governing Equations, Thermodynamics and Numerics

Subgrid-scale turbulent fluxes:


The subgrid-scale fluxes are:

ij t
qj t (2.24)
t
Jj,k

2.2.1 The filtered viscous terms


The laminar filtered stress tensor ij is given by the following relations (see [Poinsot &
Veynante 2005]):

ij = 2(Sij 13 ij Sll ),
(2.25)
2(Sij 13 ij Sll ),

and
1
ui uj
Sij = ( + ), (2.26)
2 xj xi

The filtered diusive species flux vector is:



Wk Xk c
Ji,k = Dk W xi Yk Vi
(2.27)
Dk W

k Xk
k Vi c ,
Y
W xi

where higher order correlations between the dierent variables of the expression are as-
sumed negligible.
The filtered heat flux is :

T
qi = x + Nk=1 Ji,k hk
i
(2.28)
T + N J h
xi k=1 i,k k

These forms assume that the spatial variations of molecular diusion fluxes are negli-
gible and can be modeled through simple gradient assumptions.

2.2.2 Subgrid-scale turbulent terms for LES


As highlighted above, filtering the transport equations yields a closure problem, which
requires modeling of the Subgrid-Scale (SGS) turbulent fluxes (see Eq. 2.2).
The Reynolds tensor is :

ij t = (u
i uj u
i u
j ) (2.29)

where ij t is modeled with the turbulent-viscosity hypothesis (or Boussinesqs hypothesis):



1
ij = 2 t Sij ij Sll ,
t
(2.30)
3
2.2. Filtered Equations for LES 27

which relates the SGS stresses to the filtered rate of strain, mimicking the relation between
stress and rate of strain (see Eq. 2.12). Models for the SGS turbulent viscosity t are
presented in Sec. 2.3.
The subgrid-scale diusive species flux vector is:
t


Ji,k = ui Yk u
i Yk , (2.31)

t
Ji,k is modeled with a gradient-diusion hypothesis:

t k
Wk X c,t
Ji,k = Dkt Yk Vi , (2.32)
W xi

with
t
Dkt = t
(2.33)
Sc,k

The turbulent Schmidt number Sc,k


t
= 0.6 is the same for all species. The turbulent
correction velocity reads:

N k
t W k X
Vic,t = t
, (2.34)
k=1
Sc,k W xi

with t = t /.
The subgrid-scale heat flux vector is:

qi t = (u
iE u

i E), (2.35)

where E is the total energy. The SGS turbulent heat flux qt also bears the same form as
its molecular counterpart (see Eq. 2.14):

N
T
Ji,k hk ,
t
qi t = t + (2.36)
xi k=1

with

t C p
t = . (2.37)
Prt

The turbulent Prandtl number Prt is set to a constant value of 0.6.


The correction velocity for laminar diusion then reads:

N k
c Wk X
Vi = (2.38)
k=1
Sc,k W xi .
28 Chapter 2. Governing Equations, Thermodynamics and Numerics

2.3 Models for the subgrid-stress tensor


Models for the subgrid-scale turbulent viscosity t are an essential part of a LES. The SGS
turbulence models are derived on the theoretical ground that the LES filter is spatially
and temporally invariant. Variations in the filter size due to non-uniform meshes are not
directly accounted for in the LES models. Change of cell topology is only accounted for
1/3
through the use of the local cell volume, that is = Vcell .

2.3.1 Smagorinsky model


In the Smagorinsky model, the SGS viscosity t is obtained from

t = (CS ) 2 Sij Sij
2
(2.39)

where CS is the model constant set to 0.18 but can vary between 0.1 and 0.18 depending
on the flow configuration. The Smagorinsky model [Smagorinsky 1963] was developed in
the 1960s and heavily tested for multiple flow configurations. This closure is characterized
by its globally correct prediction of kinetic energy dissipation in homogeneous isotropic
turbulence. However, it predicts non-zero turbulent viscosity levels in flow regions of pure
shear, which makes it unsuitable for many wall-bounded flows [Nicoud & Ducros 1999].
This also means that it is also too dissipative in transitioning flows [Sagaut 2002], such
as turbulent jets.

2.3.2 WALE model


In the WALE model, the expression for t takes the form:
(sdij sdij )3/2
t = (Cw ) 2
(2.40)
(Sij Sij )5/2 +(sdij sdij )5/4
with
1 1
sdij = gij 2 + g
( 2
ji ) g 2
kk ij (2.41)
2 3
Cw = 0.4929 is the model constant and gij denotes the resolved velocity gradient. The
WALE model [Nicoud & Ducros 1999] was developed for wall bounded flows and allows
to obtain correct scaling laws near the wall in turbulent flows. This model allows the
transitioning of shear flows in LES.

2.4 Real-Gas Thermodynamics


The AVBP solver first have been adapted to the real-gas framework in [Schmitt 2009]. The
thermodynamic derivations have been rephrased following [Meng & Yang 2003] and are
synthesized herein. The advantage of the current formulation is that all partial derivatives
are directly expressed as a function of transported variables, such as k = Yk instead of
molar or mass fractions.
2.4. Real-Gas Thermodynamics 29

2.4.1 Generalized Cubic Equation of State


Due to their computational eciency, cubic equation of states have been chosen for im-
plementation in the AVBP code. They allow a good overall description of supercrit-
ical fluid thermodynamics, although they are less accurate than more complex Equa-
tions Of States (EOSs) such as Benedict-Webb-Rubin (BWR) [Benedict et al. 1942].
However, the BWR EOS is mathematically uneasy to handle, and requires numerical
approximations of the partial pressure derivatives used in a CFD code. The Soave-
Redlich Kwong (SRK) [Soave 1972] and the Peng-Robinson (PR) [Peng & Robinson 1976]
EOSs are widely used in the supercritical CFD community. They are very similar and
can be written in the following generic form:

RT (T )
P = 2 (2.42)
v b v + d1 bv + d2 b2
PR : (d1 , d2 ) = (2, 1) (2.43)
SRK : (d1 , d2 ) = (1, 0) (2.44)

where P is the pressure, T the temperature, v the molar volume (v = W/), (T ) and b are
parameters computed with respect to the critical points (Tc,i , Pc,i ) of the species contained
in the mixture and their acentric factor i . In order to simplify further derivations, the
polynomial D(v) is defined as:

D(v) = v 2 + d1 bv + d2 b2 , (2.45)

and the EOS now reads:


RT (T )
P = (2.46)
v b D(v)

2.4.2 Primitive to conservative variables


The description of an initial solution for the conservation equations is generally easier
with primitive variables (T ,P ,Yk , ui ), whereas conservative variables are needed (E, k ,
, ui ). Thus, the density and energy needs to be computed from the primitive variables,
using the EOS.

Density from (T, P, Yk )


If Eq. 2.46 is multiplied by (v b) and D(v), a cubic polynome in v appears (this is
why these EOS are named cubic):

a3 v 3 + a2 v 2 + a1 v + a0 = 0 (2.47)
with a0 = b [ + (RT b) d2 b]
a1 = P d2 b2 (RT b) d1 b +
a2 = P d1 b (RT b)
a3 = P
30 Chapter 2. Governing Equations, Thermodynamics and Numerics

The roots of this polynomial are found analytically, using Cardans formula. This
formula is derived in Appendix A.1. The molar volume is determined and the density is
deduced as:
W
= (2.48)
v
The SRK and PR EOSs are used to compute the density of oxygen at 10 MPa as a
function of temperature, and are compared to the NIST database [Lemmon et al. 2009] in
Fig. 2.3. The pseudo-boiling temperature (see Eq. 1.1) is Tpb = 172 K and is identified by
a circle on Fig. 2.3. The SRK EOS is more accurate for T < Tpb whereas the PR EOS is
better for T >= Tpb . Thus, depending on the configuration, one EOS can be better suited
than the other. The PR EOS is chosen to accurately predict transcritical densities in the
DNS studies of Chap. 3. In Chap. 4, the SRK EOS is chosen to accurately predict the
injection densities far below Tpb , giving more realistic injection velocities when imposing
the mass flow rate of reactants.

1500 NIST
PR
l [kg/m ]
3

1000 SRK

500

0
100 150 200
T [K]
Figure 2.3: Density of oxygen as a function of temperature, at 10 MPa, computed with
the Peng-Robinson and the Soave-Redlich-Kwong equations of state, and compared to the
NIST database [Lemmon et al. 2009]. The circle at T = 172 K shows the pseudo-boiling
point.

2.5 CPU cost


In order to obtain converged flow statistics in DNS and LES, a sucient time period
should be simulated, which is related to the time needed for the fluid to flow through the
region of interest:
L
Tf t = (2.49)
U
2.5. CPU cost 31

with L the length of the region of interest and U a characteristic convection velocity. In
practice, 5 to 10 flow-through times are needed to obtain converged statistics.
Then, the simulation time step dt determines how many temporal iterations Nn,f t are
needed to simulate the fluid flow during n Tf t :

Tf t
Nn,f t = n (2.50)
dt
The eciency of a CFD solver can be measured with the time needed for a CPU to
compute a temporal iteration per grid point. The order of magnitude of the CPU cost
per iteration and per node of the AVBP solver is: Ci,n = 100 s.
Thus the computational cost of an unsteady numerical simulation is:

C = Nn,f t Nnodes Ci,n (2.51)

When a mesh is homogeneously refined by a factor of two in each direction, the number
of nodes increases by a factor of 2ndim and the time step decreases by a factor of 2. Thus,
the total CPU cost of a simulation increases by a factor of 2ndim +1 . This increased cost
should always be kept in mind when conducting mesh convergence studies of unsteady
simulations.
Chapter 3
A DNS study of turbulent mixing and
combustion in the near-injector region
of Liquid Rocket Engines

Contents
3.1 Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.1.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.1.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.1.3 Characteristic Numbers and Reference Scales . . . . . . . . . . . . . 37
3.1.4 Computational Grid . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1.5 Numerical Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.1.6 Initial conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.1.7 Chemical kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2 Cold Flow Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.1 Vortex Shedding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2.2 Comb-like structures . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.2.3 Scalar Dissipation Rate and turbulent mixing . . . . . . . . . . . . . 54
3.2.4 Mean flow field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
3.2.5 Influence of mesh resolution . . . . . . . . . . . . . . . . . . . . . . . 60
3.3 Reacting Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3.1 Flame stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3.2 Vortex shedding and comb-like structures . . . . . . . . . . . . . . . 64
3.3.3 Combustion Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.3.4 Mean flow field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.3.5 Comparison of numerical results with existing experimental data . . 73
3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

In the presence of high convection speeds inside the combustion chamber of a Liquid
Rocket Engine (LRE), it is crucial that the flames stay anchored at the injector rim,
to prevent blow-o and extinction. Understanding the flame stabilization mechanisms
Chapter 3. A DNS study of turbulent mixing and combustion in the
34 near-injector region of Liquid Rocket Engines

is thus important for the design of rocket engines. The literature on flame stabilization
is extremely large, with detailed investigations using both experiments and numerical
simulations. The main parameter characterizing flame stabilization is the distance from
the injector to the flame front, and a wide variety of phenomena have been reported with
chemistry, transport and various sources of heat losses, possibly playing an important
role.
In the pioneering numerical work of [Oefelein & Yang 1998,Oefelein 2001,Oefelein 2005,
Oefelein 2006], important features of the flame stabilization mechanism in the context of
LREs have been identified. For instance, it has been shown that the unsteady stagna-
tion point behind the lip of a coaxial injection contains recirculated fuel-rich combustion
products that supply heat to the flame tip and help stabilization.
However, there are still uncertainties on the flame stabilization mechanisms. Indeed,
experimental studies of transcritical jet flames [Singla et al. 2005, Singla et al. 2007] have
shown that the flame is stabilized at a small but finite distance from the injector rim,
which indicates the presence of partial premixing. It is argued in [Oefelein 2005] that heat
loss from the flame to the injector wall could provoke flame quenching in the experimental
studies, which is not taken into account in the numerical simulations.
In addition, turbulent mixing processes in supercritical flows are not well known and
understood, in particular, because their experimental observation is not easy. Therefore,
it is proposed here to conduct a DNS of turbulent flame stabilization in conditions of
LRE. The main objective is to clarify the discrepancies on flame stabilization distance
mentioned above. Mixing and combustion regimes will be scrutinized in fully resolved
simulations and driving mechanisms will be determined. This will also help to identify
paths for improvement of turbulent mixing and combustion models for LES.

3.1 Configuration
The configuration chosen for the DNS study corresponds to a 2D cut through a LOx/GH2
coaxial injector, representative of an injector of a LRE:
the ambient pressure is high: P = 10 MPa, about twice as much as the critical point
of O2 (5.04 MPa)

the inner dense oxygen jet has a low velocity

the outer light hydrogen jet has a large velocity. The large shear, induced by the
velocity dierence between the jets, enhances strong turbulent mixing.

the dimensions of the injector are small (a few millimeters), with a lip height of
h = 0.5 mm.

the inner O2 tube is recessed from the injection plane, to enhance turbulent mixing.
A sketch of the computational domain is displayed in Figs. 3.1 a) and b).
This configuration is inspired from the previous numerical work of [Oefelein & Yang 1998,
Oefelein 2005]. The chosen thermodynamic conditions (transcritical injection) are also
3.1. Configuration 35

Figure 3.1: a) Typical coaxial injector of a LRE. b) Boundary conditions for the 2D
computational domain

close to the experimental study of transcritical LOx/LCH4 combustion from [Singla et al. 2007],
and will enable a qualitative comparison with experimental data in Sec. 3.3.5. In [Zong
& Yang 2006, Zong & Yang 2007], a supercritical LOx/LCH4 coaxial flow with the same
injection conditions as [Singla et al. 2007] has been studied. The present results will also
be compared with this numerical work.

3.1.1 Thermodynamics

Cold flow case


Real-gas thermodynamics is accounted for through the PR EOS [Peng & Robinson 1976]
(see Sec. 2.4.1) while transport coecients are modeled based on the theory of correspond-
ing states for the dynamic viscosity and the thermal conductivity (see Sec. 2.1.4) and con-
stant Schmidt numbers (see Tab. 3.1). This overall numerical and modeling methodology
has already been validated for non reacting flows [Schmitt et al. 2010b].

Reacting flow case


The critical-point coordinates of the combustion intermediate species OH, O, H, H2 O2
and HO2 , for which no experimental data is available, is estimated using the Lennard-
Jones potential-well depth i , and the molecular diameter i of the i-th species from the
CHEMKIN transport coecients of the San Diego Mechanism, according to the following
Chapter 3. A DNS study of turbulent mixing and combustion in the
36 near-injector region of Liquid Rocket Engines

expression [Giovangigli et al. 2011]:


Vc,i = 3.29N i3 (3.1)
i
Tc,i = 1.316 (3.2)
k
(3.3)
where N is the Avogadro number and k is the Boltzmann constant. The acentric factor
ac is set to zero for radical species. The numerical values for all species are summarized
in Tab. 3.1.
Parameters H2 O2 H2 O O H OH H2 O 2 HO2
Tc,i (K) 33 154.581 647.096 105.28 190.82 105.28 141.34 141.34
Pc,i (MPa) 1.2838 5.0430 22.064 7.0882 31.013 7.0883 4.7861 4.7861
Vc,i (cm /mol)
3
64.284 73.368 55.948 41.205 17.069 41.205 81.926 81.926
ac -0.216 0.0222 0.3443 0.0 0.0 0.0 0.0 0.0
Schmidt Number 0.28 0.99 0.77 0.64 0.17 0.65 0.65 0.65

Table 3.1: Species critical-point properties (temperature T , pressure P , molar volume V


and acentric factor ) and Schmidt numbers.

3.1.2 Boundary Conditions


The boundary conditions are represented on Fig. 3.1 b). A 1/7th power law for the inlet
boundary condition is used to mimic the mean turbulent velocity profile in a pipe flow:
1/7
y ywall
u(y) = U inj
(3.4)
4.5h
with U inj , the bulk injection velocity (see Tab. 3.2) and y ywall , the distance to the
injector wall.
The outlet pressure is imposed at 10 MPa, using the NSCBC formalism [Poinsot
& Lele 1992, Baum et al. 1994] and accounts for both real-gas eects [Okongo & Bel-
lan 2002a] and transverse terms [Granet et al. 2010].
A 1h-long sponge layer is placed at the exit of the computational domain, to prevent
spurious acoustic wave generation due to the outgoing hydrodynamic structures. This
makes the computational domain very compact (10h 11h) and results in a moderate
CPU cost for the simulations (see Sec. 3.1.4).
The upper and lower boundaries are slipping walls while the splitter plate is an adia-
batic no-slip wall.
No synthetic turbulence is added to the inflow boundary conditions, and yet, strong
turbulence levels caused by vortex shedding are observed downstream the lip. This allows
the development of a turbulent mixing layer and strong flame/turbulence interactions
in the reacting case. It is believed that this 2D phenomenon (vortex shedding at the
lip) drives the flame turbulence interaction, which justifies the 2D nature of the present
computations.
3.1. Configuration 37

3.1.3 Characteristic Numbers and Reference Scales


The main scales and non-dimensional numbers characterizing the flow are summarized in
Tab. 3.2, where R and R (defined in Eqs. 1.9 and 1.10) are the reduced temperature
and pressure, respectively.

T [K] inj [kg/m3 ] Uinj [m/s] [Pa.s] R R Re Ma


O2 100 1258 30 3.9 104 0.65 2.0 9 105 0.04
H2 150 15.8 125 5.1 106 4.55 7.8 2 105 0.12
R Ru J
80 4 0.2
Uconv [m/s] conv [s] Tf t [s]
39.6 12.5 125

Table 3.2: Characteristic flow quantities

A transcritical injection has been chosen (R < 1 at injection) to mimic realistic the
injection conditions of a LRE. The resulting density ratio R (see Eq. 1.14) is very large
and plays an important role in the turbulent mixing characteristics.
The PR EOS is chosen to accurately predict densities around and above the pseudo-
boiling temperatures (see Fig. 2.3) that are encountered in the reacting case.
The momentum-flux ratio J, defined in Eq. 1.15, is much lower than typical LRE val-
ues, where J lies between 1 and 20. This stems from the high velocity of oxygen, which
would be lower in a real LRE. However, this change of velocity lowers the computational
cost to reach statistical convergence, by lowering the flow-through time, without dramat-
ically changing the characteristic numbers. Indeed, with a more realistic momentum-flux
ratio of 2, the oxygen velocity would be 3 m/s, which would result in a low Mach number
of 0.004 and a large Reynolds number of 9 104 .
The Mach number is rather small in both the LOx and GH2 stream. Even in the
reacting case, the Mach number stays below 0.2 so that compressibility eects are small.
The convective velocity of the coherent structures in the flow can be evaluated using
the expression given in [Papamoschou & Roshko 1988] and also used for instance in
[Raynal 1997]:
1/2
H2
UOinj
2
+ O2
UHinj2
Uconv = 1/2 (3.5)
H 2
1+ O2

This formula is derived with the assumption that in the moving reference frame of the shed
vortices, a stagnation point exists between the two streams. Thus, using the Bernoulli
formula, with equal dynamic pressures at the stagnation point, the above relation is
deduced. This relation appears to be valid for the low Mach number considered herein,
where compressibility eects are negligible.
Chapter 3. A DNS study of turbulent mixing and combustion in the
38 near-injector region of Liquid Rocket Engines

This convective velocity, is used in conjunction with the splitter plate height h, to
define a convective time:
h
conv = (3.6)
Uconv

3.1.4 Computational Grid


In the cold flow case, the main phenomenon to capture is the turbulent mixing of the
two separated streams behind the lip, due to vortex shedding in the lip wake. In the
hot flow case, the vortex shedding imposes a large and variable strain on the flame and
the computational domain was successively refined in order to obtain at least 5 points in
the radical species profile in the most stretched regions of the flame, ensuring a proper
discretization of both the flame and the turbulence, leading to a mesh resolution of =
h/500 (= 1 m) This fine resolution is applied in a layer containing the lip wake, over a
3h vertical extent. Outside this zone, a transverse stretching factor of 1.02 is employed.
The mesh finally contains 13.5 million nodes and quadrilaterals. As a comparison with
previous eorts to compute such configurations [Oefelein 2005], the resolution is ten times
greater.
The integral length scale is of the order of lt = h and the velocity fluctuations in
the cold flow case are of the order of u = 0.3 Us (see Fig. 3.23). Using the kinematic
viscosity of the hydrogen stream and assuming a fully developed 3D turbulent flow, the
scale separation is:
3/4
= lt Ret (3.7)
6
= 0.16 10 m
Thus, dx/6. Since the present computation is 2D, there is no theoretical expression
for scale separation, and the expression of scale separation is only used to assess spatial
resolution requirements. Due to the high precision of the current numerical scheme, the
mesh spacing does not need to be equal to the Kolmogorov scale but simply needs to
be O() [Moin & Mahesh 1998], which is the case here. It is also checked a posteriori
in Fig. 3.16 that no energy packs up at large wave numbers, which indicates that either
turbulence has not yet developed, or that it has been dissipated by viscous forces at small
resolved scales.
In order to compare cold and hot flow simulations, the same mesh was used for both
cases.

Probe location
Fig. 3.2 shows the location of probes placed in the computational domain, which enable
a time resolved analysis of all turbulent scales. The probe arrangement is a grid within
the mixing layer.

3.1.5 Numerical Scheme


The compressible Navier-Stokes equations are solved using the AVBP code presented in
Sec. 2. The TTG4A scheme is used in the cold and hot flow cases. This scheme is third
3.1. Configuration 39

Figure 3.2: Probes located in the mixing layer.

order accurate in space and time. The transcritical density gradient requires numerical
stabilization and artificial dissipation is added in this region, using the real-gas operator
described in [Schmitt 2009]. The mesh convergence study conducted in Sec. 3.2.5 allows
to verify a posteriori that numerical stabilization of the transcritical density gradient does
not have an impact on the time-average solution and RMS fluctuations.

CPU cost
Using the length of the mixing layer L = 10 h in Eq. 2.49, the flow-through time of
coherent structures is Tf t = 125 s (using Eq. 3.5).
In the non-reacting case and the reacting case, the time step is approximately dt =
0.5 109 s and dt = 0.3 109 s, respectively. A total of 2 106 temporal iterations were
simulated in both cases, which corresponds to 8 Tf t and 5 Tf t for the non-reacting case
and reacting case, respectively (see Eq. 2.50). These simulations have been conducted on
the JADE supercomputer at CINES (Intel Quad-Core at 2.8 GHz), where the eciency of
AVBP is Ci,n = 15 s for the non-reacting case (2 species are transported) and Ci,n = 45 s
for the reacting case (8 reacting species are transported). Thus, the simulation cost is
approximately 100 103 CPU hours and 300 103 CPU hours for the non-reacting and
reacting case, respectively.

3.1.6 Initial conditions


Both initial conditions for the non-reacting and reacting cases are described in this section.
Chapter 3. A DNS study of turbulent mixing and combustion in the
40 near-injector region of Liquid Rocket Engines

Cold flow case


The initial solution is chosen to minimize the convergence time, using the boundary
conditions swept through the whole domain. A hyperbolic tangent profile for the O2
mass fraction and for the temperature is used behind the splitter plate, along with a zero
velocity zone, as shown in Fig. 3.3. The thickness of the hyperbolic tangent profile is
= h/4, so that the oxygen and temperature gradient thicknesses are initially resolved
on more than 100 grid points, and yet are steep enough to trigger the flow transition.

0.8
YO
0.6 2

u/u
0.4 H
2

(TT )/(T T )
0.2 O H O
2 2 2

0
5 3 1 1 3 5
y/h [ ]
Figure 3.3: Transverse cut through the initial solution of the cold flow, downstream the
lip.

Reacting flow case


Combustion is initiated in the computational domain with the following procedure:
1. first, the field of mixture fraction is computed from a non-reacting established flow,

2. then, the flame structure from a laminar diusion flame is applied onto this field
with subsequent replacement of the composition and temperature.
During this procedure, the cold-flow pressure and velocity fields are not altered. As
expected, significant acoustic perturbations are generated when the flow adapts from the
approximated to the exact solution of the conservation equations. However, the procedure
is successful because the pressure waves eventually leave the computational domain and
a stable combustion regime is reached, as shown in Fig. 3.4.

3.1.7 Chemical kinetics


The combustion of hydrogen and oxygen is modeled using a detailed scheme accounting
for 8 species and 12 reactions [Boivin et al. 2011], which is derived from the San Diego
3.1. Configuration 41

7
x 10
1.5
P [Pa]
1

0.5
2 4t/ [ ]6 8 10
conv

Figure 3.4: Temporal evolution of the minimum, mean and maximum pressures in the
computational domain, after instantaneous ignition of the cold flow case.

mechanism [Petrova & Williams 2006]. The forward rate coecients are given in Tab. 3.3.
The backward reaction rates are classically computed using low-pressure entropy and
enthalpy NIST/JANAF tables [Lemmon et al. 1998]. A more general treatment of the

Reaction Aa n Ea
1 H+O2 OH+O 3.52 1016 -0.7 71.42
2 H2 +O OH+H 5.06 104 2.67 26.32
3 H2 +OH H2 O+H 1.17 109 1.3 15.21
4 H+O2 +M HO2 +Mb 4.65 1012 0.44 0.0
5 HO2 +H 2OH 7.08 1013 0.0 1.23
6 HO2 +H H2 +O2 1.66 1013 0.0 3.44
7 HO2 +OH H2 O+O2 2.89 1013 0.0 2.08
8 H+OH+M H2 O+Mc 4.00 1022 -2.0 0.0
9 2H+M H2 +Mc 1.30 1018 -1.0 0.0
10 2HO2 H2 O2 +O2 3.02 1012 0.0 5.8
11 HO2 +H2 H2 O2 +H 1.62 1011 0.61 100.14
12 H2 O2 +M 2OH+Md 2.62 1019 -1.39 214.74

Table 3.3: Forward rate coecients in Arrhenius form k = AT n exp (E/RT ) for the
skeletal mechanism.
Units are mol, s, cm3 , kJ, and K.
a
b
Chaperon eciencies are 2.5 for H2 , 16.0 for H2 O and 1.0 for all other species.
c
Chaperon eciencies are 2.5 for H2 , 12.0 for H2 O and 1.0 for all other species.
d
Chaperon eciencies are 2.0 for H2 , 6.0 for H2 O and 1.0 for all other species.

reaction rates in the real-gas framework, using chemical potentials directly computed
from the equation of state has been investigated by [Giovangigli et al. 2011]. However,
the impact on the flame structure appears to be rather small and tends to indicate that
the perfect-gas treatment of the reaction rates is a good approximation. This is mainly
due to high temperatures and relatively low pressures (and hence perfect-gas behavior)
at the flame location. Note that since high pressure is considered in the present study,
Chapter 3. A DNS study of turbulent mixing and combustion in the
42 near-injector region of Liquid Rocket Engines

the high-pressure limit of the fallo reactions have been taken.


The validation of the implementation in AVBP is achieved by comparing the flame
structure using CANTERA [Goodwin 2002] and AVBP in a counterflow flame configura-
tion. The numerical setup is that of an opposed-jet flame [Pons et al. 2009] computed, in
AVBP, on a simple square mesh of size h with a constant grid size identical to that of the
splitter case. The left boundary condition is a symmetry, the right side an outlet, while
hydrogen comes from the top and oxygen from the bottom. The boundary velocity are
determined to impose a constant strain, a, on the flame:

uO2 (x) = ax (3.8)


vO2 (y) = ay (3.9)
1
O2 2
uH2 (x) = ax (3.10)
H 2
1
O2 2
vH2 (y) = ay (3.11)
H 2
(3.12)

The hydrogen velocity is chosen so that the momentum flux u2 is the same between the
oxygen and hydrogen stream, to place the stagnation point in the middle of the computa-
tional domain. As for the value of the strain rate, a, the validation is conducted at a value
based on the splitter height h and the mean velocity dierence between the two streams
at the injection, which yields a = 3800 s1 . A typical result of such strained diusion
flame is presented in Fig. 3.5 showing streamlines superimposed on the temperature field.

Figure 3.5: Strained diusion flame computed in AVBP: streamlines superimposed on the
temperature field. Thermodynamic conditions correspond to the splitter case: hydrogen
at 150 K from the top, oxygen at 100 K from the bottom and ambient pressure is 10 MPa.

The validation procedure comprises three computations:


3.1. Configuration 43

1. CANT_PG: a computation is performed using CANTERA using the perfect-gas


equation of state. The temperature of the fresh gases is 300 K to limit real-gas
eects.

2. AVBP_PG: this is the same computation as CANT_PG, performed in AVBP.


Because CANTERA solves for the full transport matrices while AVBP assumes
constant Schmidt numbers and mixture-averaged transport coecients, this com-
putation is both a validation of the implementation of the chemistry in AVBP as
well as a validation of the simplified transport.

3. AVBP_RG: for this computation, the temperature of the fresh gases is lowered to
match those of the splitter case and evaluate real-gas eects on the flame structure.

A cut through the flame at x = h/2 in the computational domain of AVBP is compared to
the flame structure from CANTERA. Figure 3.6 shows the temperature and mass fraction
of HO2 (the main initiator of the combustion) versus the mixture fraction (see Eq. 1.3)
for the three computations.

(a) (b)

Figure 3.6: Comparison of flame structure versus mixture fraction between AVBP and
CANTERA. (a) temperature and (b) HO2 mass fraction. The vertical bar indicates the
stoichiometric mixture fraction.

First comparing simulations CANT_PG and AVBP_PG, the agreement between the
two codes is excellent: the maximum discrepancy for the temperature is of the order of
70 K, which gives a relative error of 2 %. The dierences for the mass fraction YHO2
are even smaller. This agreement validates the implementation in AVBP as well as the
assumptions on the transport, for this temperature. The laminar flame in the ther-
modynamic conditions of the splitter case (AVBP_RG) is very similar to the higher-
temperature computation (AVBP_PG) in the flame region. This conclusion is similar to
that of [Ribert et al. 2008], justified by the fact that the combustion processes take place
in hot regions where real-gas eects are negligible.
It is important to note, for later analysis of flame structures, that far from extinction,
the structure of a non-premixed flame at a large Damkhler number is very close to the
equilibrium lines, in the mixture fraction space. This is shown in Fig. 3.7, where the
major species mass fractions are plotted against mixture fraction. It will thus be dicult
to distinguish between a pure diusion flame and a partially-premixed flame, solely looking
Chapter 3. A DNS study of turbulent mixing and combustion in the
44 near-injector region of Liquid Rocket Engines

at the flame structure and local analysis in the physical space will have to be performed.

H2
1
O2
H2O
Y[]

0.5

0
0 0.2 0.4 0.6 0.8
Mixture fraction [ ]

Figure 3.7: Major species mass fractions as a function of mixture fraction for the non-
premixed counterflow flame configuration (lines) and equilibrium (symbols)

3.2 Cold Flow Results


In this section, the non-reacting flow is analyzed with specific attention to the interaction
between turbulence and the large density gradients.

3.2.1 Vortex Shedding


Figure 3.8 shows an instantaneous axial velocity field. Small vortices are regularly shed
from to the top corner of the lip, due to the shear-induced Kelvin-Helmholtz instability.

These vortices are either transported downstream, or trapped into the recirculation
zone behind the lip. In both cases, many pairing events occur, as exemplified in Fig. 3.9,
which ultimately forms large vortices having a length scale comparable to the splitter
height. The dash lines follow the displacement of the small developing vortices, and
allows to determine a visual convection speed of 5 m/s. This convection speed is smaller
than the convection speed of the largest coherent vortices because, in the vicinity of the
injector wall, the velocity magnitude is small (see Eq. 3.4). On the first four images of
Fig. 3.9, the distance between vortices is approximately h/8. Using the convection speed
of 5 m/s, the corresponding Strouhal number is equal to 1.0. Because of vortex merging,
the distance between the vortices on the last four images of Fig. 3.9 is increased and equals
3h/8, which corresponds to a Strouhal number of 0.3. Thus, subharmonics are created by
vortex merging, which should be visible on the spectrum of the probe signals placed in
the mixing layer. The vortex observed at the center of Fig. 3.8 results from this merging
process.
Further downstream, the largest coherent vortices break down into smaller fluctuating
motions, as observed at the right side of Fig. 3.8, while the flow transitions towards
fully-developed turbulence.
3.2. Cold Flow Results 45

Figure 3.8: Non-reacting flow: axial velocity field, showing the shear-induced Kelvin-
Helmholtz instability.

This large range of turbulent scales rapidly mixes the dense oxygen with the light
hydrogen, as can be seen on the oxygen mass fraction field shown in Fig. 3.10.
Figs. 3.11 and 3.12 show a temporal evolution of the transverse velocity and the oxygen
mass-fraction fields, respectively. The time interval between two snapshots is conv and
the dash lines follow the displacements of the largest coherent structures. These lines
allow to determine a visual convection speed of the largest coherent structures Uconv,m
equal to 31 m/s. This value is slightly lower than the value of 39 m/s found using Eq. 3.5.
However, using the phase velocity derived in [Juniper 2001]

H2 inj
UOinj
2
+ U
O2 H2
Up = , (3.13)
1 + HO2
2

one finds exactly 31 m/s. In [Juniper 2001], it is conjectured that the convection speed
of the coherent structures might start at Up , given by Eq. 3.13 and tends towards Uconv ,
given by Eq. 3.5.
The wavelength of the largest coherent structures is approximately 5 h (distance be-
tween two dash lines). The time scale of the largest coherent structures is thus =
5h/Uconv,m 6.5 conv , which correspond to a frequency of f = 12 kHz. The Strouhal
Chapter 3. A DNS study of turbulent mixing and combustion in the
46 near-injector region of Liquid Rocket Engines

1 5

2 6

3 7

4 8
h 2h 3h h 2h 3h
4 4 4 4 4 4
1
Figure 3.9: Non-reacting flow: oxygen mass fraction field at the H2 corner of the lip, at
successive instants. The time interval between frames is 10 s

Figure 3.10: Non-reacting flow: O2 mass fraction field, showing the rapid mixing of the
two streams by a large range of vortical structures.

number, which is a non-dimensional frequency, is defined with the convective time:

St = f conv (3.14)

The Strouhal number of the largest coherent structure, assessed visually, is thus St = 0.15.
3.2. Cold Flow Results 47

Fourier transform
First, a single probe signal within the mixing region immediately before the outlet
boundary condition (before the sponge layer, at i=10 and j=12 in Fig. 3.2) is plotted
in the temporal and Fourier space, in Fig. 3.13. Although it is clear that a frequency
dominates the signal (St = 0.14, the dash line plotted with the temporal signal), many
secondary peaks are also observable. Note that the plotted power spectrum is actually a
random variable and the average periodogram procedure [Oppenheim et al. 1989] must
be used to reduce the variance of the spectrum estimate. Due to the small amount of
time simulated, the Welch averaging procedure greatly deteriorates frequency resolution
as shown in Fig. 3.14, for various numbers of windows. Since the total time simulated is
80 conv , the strouhal number resolution is of the order of 102 , without spectrum aver-
aging. Using 8 Welch windows approximately divides this resolution by 10, as shown in
Fig. 3.14, but still enables the identification of a deterministic vortex shedding frequency:
St = 0.18. This frequency is equal to the one determined visually, using Fig. 3.11.
The Fourier transform is then applied to the signals of the probes located at j = 17,
between i = 1 and i = 12 on Fig. 3.2, which are located on the passage of the vortices shed
from the H2 corner of the lip. The axial evolution of the transverse velocity power density
spectrum is plotted on Fig. 3.15. This figure shows that the vortex shedding strouhal
number of the largest coherent structures (energy containing eddies) is St 0.2, which
is consistent with the previous visual observations and which is in the range of vortex
shedding strouhal numbers observed in the range of 0.1 to 0.3 in multiple experimental
studies at large Reynolds numbers (see [Roshko 1961] and [Williamson & Brown 1998],
among others). In [Zong & Yang 2006, Zong & Yang 2007], a supercritical LOx/LCH4
coaxial flow has been studied, and a frequency approximately equal to 0.2 was also found
for vortex shedding, which is consistent with the present result. There are persisting
strouhal number bands in Fig. 3.15 for 1 < St < 5, which shows the existence of smaller
coherent structures (subharmonics), as already observed in Fig. 3.9.
The power spectrum density of the squared transverse velocity (E22 = u2 u2 , where
Eij is the Reynolds stress tensor) located at i=10, j=12 is shown in logarithmic scale in
Fig. 3.16. For fully developed 2D turbulent flows, the spectrum inertial range exhibits a
3 slope [Rutgers 1998, Vallgren & Lindborg 2011]. A similar slope is observed in the
present numerical simulation, which tends to indicate that the 2D turbulence is nearly
developed at this location.
Chapter 3. A DNS study of turbulent mixing and combustion in the
48 near-injector region of Liquid Rocket Engines

Figure 3.11: Non-reacting flow: temporal evolution of the transverse velocity field. The
time interval between two snapshots is conv
3.2. Cold Flow Results 49

Figure 3.12: Non-reacting flow: temporal evolution of the oxygen mass-fraction field. The
time interval between two snapshots is conv
Chapter 3. A DNS study of turbulent mixing and combustion in the
50 near-injector region of Liquid Rocket Engines

100

50
v [m/s]




0 5 10 15 20 25 30 35 40
WW0)/oconv

0.03
PSD [m.s /Hz]

0.02
2 

0.01

0  0 1
10 0.14 10 10
St = f o
conv

Figure 3.13: Non-reacting flow: transverse velocity signal analysis. Top) temporal
signal; dominant harmonic (St = 0.14). Bottom) Power Spectrum Density without
spectrum averaging.

0.035

0.03
nw = 2
nw = 4
0.025 nw = 8
PSD [m2.s/Hz]

0.02

0.015

0.01

0.005

0  0 1
10 0.14 0.18 10 10
St = f oconv

Figure 3.14: Non-reacting flow: eect of the Welch averaging procedure on the spectrum
frequency resolution. Number of windows=2,4,8.
3.2. Cold Flow Results 51

Figure 3.15: Non-reacting flow: spatial evolution of the transverse velocity spectrum in
the wake of the lip. Eight Welch averaging windows are used.
Chapter 3. A DNS study of turbulent mixing and combustion in the
52 near-injector region of Liquid Rocket Engines

2
10
PSD [m4.s4/Hz]

1
10
0
10
1
10
2 3 slope
10
3
10
4
10 2 1 0 1
10 10 10 10
St
Figure 3.16: Non-reacting flow: power spectrum density of the squared transverse velocity
at i=10, j=12. Eight Welch averaging windows are used.
3.2. Cold Flow Results 53

3.2.2 Comb-like structures


Figure 3.17 shows the vorticity field superimposed on the high-density region (fluid regions
with a density that is higher than 0.5 = 0.5 (inj inj
H2 + O2 ) are painted in black.). The
emergence of low-speed oxygen fingers inside the high-speed hydrogen stream enhances
shear-induced instability and creates small vortices. These small vortices increase the
exchange surface between reactants, stretch the material interface and also feed large
vortices through merging, which maintain the rise of the oxygen fingers. These finger-
like structures are visually similar to experimental observations of transcritical mixing of
coaxial jets (see Sec. 1.2.2, Fig. 1.6). In the absence of surface tension, the finger-like
structures do not form droplets and are therefore further broken down by turbulence.

Figure 3.17: Non-reacting flow: vorticity field superimposed on the high-density region
(fluid regions with a density that is higher than 0.5 = 0.5 (inj inj
H2 + O2 ) are painted in
black.).
Chapter 3. A DNS study of turbulent mixing and combustion in the
54 near-injector region of Liquid Rocket Engines

3.2.3 Scalar Dissipation Rate and turbulent mixing


The scalar dissipation rate , defined in Eq. 1.2, is a measure of the molecular mixing
between hydrogen and oxygen. This mixing is favorable for combustion and is greatly
enhanced by turbulence: the wrinkled exchange surface enables a larger total mass flux
and strain locally steepens the composition gradients, which is also favorable for species
diusion.
Figure 3.18 shows a snapshot of the O2 mass fraction, along with a grey isocontour
of mid-density ( = 637 kg.m3 ) and a white isocontour of high scalar dissipation rate
( = 5 103 s1 ). This figure suggests that the majority of the scalar dissipation occurs
on the light side of the fluid. The large molecular flux of oxygen leaving the dense LOx
stream shifts the molecular mixing zone away from the transcritical density gradient. This
figure is also a great example of the turbulence enhancement of mixing: the small vortices
shed at the H2 corner stretch the LOx/GH2 material interface and merge to form larger
vortices that roll-up the O2 stream and traps pockets of H2 in between two comb-like
structures. At the right part of the figure, a dash circle shows a region where the scalar
dissipation rate is low because mixing has already occurred.

Figure 3.18: Non-reacting flow: O2 mass fraction, along with a grey isocontour of mid-
density ( = 637 kg.m3 ) and a white isocontour of high scalar dissipation rate ( =
5 103 s1 ). The dash circle shows a region of low scalar dissipation rate where mixing has
already occurred.

To confirm that scalar dissipation rate is larger in the light part of the fluid, Fig. 3.19
shows the mean and maximum values of , conditioned on mixture fraction. It is notable
that the maximum of scalar dissipation rate in the dense part of the fluid (below the
stoichiometric mixture fraction Zst = 1/9) is several orders of magnitude lower than in
the light part of the fluid (max 105 s1 , at Z = 0.6).
The scalar dissipation rate at the stoichiometric mixture fraction is of primary interest
here, because it determines the order of magnitude of the flame thermal dissipation in the
reacting case. At stoichiometry, the maximum scalar dissipation rate st,max = 6 103 s1
3.2. Cold Flow Results 55

6
10 Dense Zst
5
Light
10
4
10
r [1/s]
3 Mean
10
 Max
10
1
10
0
10   0
10 10 10
Mixture fraction [ ]
Figure 3.19: Non-reacting flow: conditioned average and maximum value of scalar dissi-
pation rate as a function of mixture fraction.

is several orders of magnitude below the extinction scalar dissipation rate, which is ap-
proximately 3 106 s1 [Kim et al. 2011]. The scalar dissipation rate at stoichiometry for
the counterflow flame used for validation in Fig. 3.6 is 104 s1 and is comparable to the
maximum value encountered in this cold flow case.
Chapter 3. A DNS study of turbulent mixing and combustion in the
56 near-injector region of Liquid Rocket Engines

3.2.4 Mean flow field


The observation of time-averaged solutions enables to quantify turbulent mixing.
Average solutions have been obtained sampling every 500 time steps ( conv /50), over
a time period of approximately 8 Tf t (= 80 conv ) in the non-reacting case and 5 Tf t in
the reacting case (see Sec. 3.1.4). In the non-reacting case, the statistical results stay
qualitatively the same between 5 Tf t and 8 Tf t (not shown), thus only 5 Tf t are simulated
in the reacting case.
Figure 3.20 shows the resulting average density field, using a logarithmic scale. On

Figure 3.20: Non-reacting flow: average density field

Fig. 3.20, three iso-values of density are plotted at 0.1 , 0.5 and 0.9 , where is defined
as
= inj inj inj
H2 + (O2 H2 ) (3.15)
The location of the isoline at may be determined as y, (x), defined by:
(x, y, (x)) = (3.16)
The spreading angles of the LOx stream may then be calculated as:
, = tan1 (y, /x). (3.17)
Fig. 3.21 shows the spreading angles for the 3 values of of Fig. 3.20.
The spreading angles converge towards a constant value before the end of the compu-
tational domain, which tends to indicate a self-similarity state is reached. The 0.1 isoline
indicates the opening of the oxygen jet, and the spreading value of +5 obtained in the
numerical simulation is close to the value of 6 degrees found over a large range of density
ratio in low pressure jets [Chen & Rodi 1980] and also well within the experimental scatter
of transcritical injection experiments (see Fig. 14.b in [Schmitt et al. 2010b]). The 0.9
isoline allows to determine a possible dense core length, although 2D turbulence is not
the same as 3D turbulence, which is defined as:
L = R/tan(,0.9 ). (3.18)
3.2. Cold Flow Results 57

With R = 4.5 h (radius of the oxygen injector) and ,0.9 = 5 , L 10R, which again
is within the 10R to 20R range found in transcritical injection experiments (see Fig. 12b
in [Oschwald et al. 2006]).

Figure 3.21: Non-reacting flow: spreading angles , .

Transverse cuts through the average fields are plotted on Figs. 3.22, 3.23 and 3.24.
Turbulent intensity reaches high values of about 30%, which is well above a typical 10%
turbulent intensity in jet developed turbulence, because of the large coherent structures
developing in the mixing layer. From the RMS plots of the axial and transverse velocity,
one can see that the lower and upper boundary conditions are actually confining the
jets, since uRM S do not reach zero and since the vRM S profile is obviously influenced
by the upper boundary condition. This confinement influences the jet dynamics but is
representative of a coaxial injector with a recessed LOx tube, as explained in Sec. 3.1.
The RMS velocities are normalized by the velocity dierence between the two streams:

Us = UHinj2 UOinj
2
. (3.19)

This correspond to the maximum fluctuation velocity for turbulent motions in a non-
reacting flow. In terms of intensity, both components of the velocity reach similar values
( 30 %) that are already attained at x = h. This highlights again the intense turbulent
activity behind the splitter. As vortical structures are transported and grow downstream
the splitter plate, fluctuations spread, with deeper penetration into the light hydrogen
stream, which is consistent with the temporal mixing layer study of [Okongo & Bel-
lan 2002b].
Chapter 3. A DNS study of turbulent mixing and combustion in the
58 near-injector region of Liquid Rocket Engines

0h
2h
Density [kg/m3]

1000
4h
6h
500

0
2 1 0 1 2
y/h
Figure 3.22: Non-reacting flow: transverse cuts of the density at 4 axial positions between
x=0h and x=6h.

0.3
1h
5h
0.2
uRMS/Us

0.1

0
5 4 3 2 1 0 1 2 3 4 5
y
Figure 3.23: Non-reacting flow: transverse cuts of the RMS axial velocity normalized by
the velocity dierence Us = UHinj2 UOinj
2
at x=1h and x=5h.
3.2. Cold Flow Results 59

0.3
1h
vRMS/Us

0.2 5h

0.1

0
5 4 3 2 1 0 1 2 3 4 5
y
Figure 3.24: Non-reacting flow: transverse cuts of the RMS transverse velocity normalized
by the velocity dierence Us = UHinj2 UOinj
2
at x=1h and x=5h.
Chapter 3. A DNS study of turbulent mixing and combustion in the
60 near-injector region of Liquid Rocket Engines

3.2.5 Influence of mesh resolution


In order to verify mesh convergence, a simulation with a coarser mesh has been conducted.
With this coarse mesh, the splitter height h contains 100 grid points (instead of 500). In
the following, the coarse and refined mesh are called h100 and h500, respectively.

Instantaneous visualization
Figure 3.25 shows instantaneous snapshots of the transverse velocity for the h100 and
h500 meshes. Even though the smallest scales are missing from the h100 mesh, the flow
fields are qualitatively similar, with small vortices shed at the H2 corner of the lip, which
merge further downstream and roll-up the oxygen stream.

(a)

(b)

Figure 3.25: Instantaneous snapshots of the transverse velocity and the density for the
(a) h100 and (b) h500 meshes.

Time-average solutions
Figure 3.26 and Fig. 3.27(a) show the mean and RMS profiles for the h100 and h500
meshes at x = 5h, which is located in between the injection and the outlet plane. The
agreement between the h100 and h500 meshes for both the RMS and mean profiles is very
good, which proves that mesh convergence is attained for the results presented in Sec. 3.2,
in spite of the numerical stabilization of the transcritical density gradient.
3.2. Cold Flow Results 61

u [m/s] 110

70 h100
h500
30

5 3 1 1 3 5
y/h [ ]
(a)

6
4
v [m/s]

2 h100
0 h500
2
4
5 3 1 1 3 5
y/h [ ]
(b)

1
0.8
YO [ ]

0.6 h100
2

0.4 h500
0.2
0
5 3 1 1 3 5
y/h [ ]
(c)

Figure 3.26: Mean profiles, at x = 5h for the h100 and h500 meshes of (a) axial velocity,
(b) transverse velocity and (c) O2 mass fraction
Chapter 3. A DNS study of turbulent mixing and combustion in the
62 near-injector region of Liquid Rocket Engines

30
24
uRMS [m/s]

18 h100
12 h500
6
0
5 3 1 1 3 5
y/h [ ]
(a)

50
40
vRMS [m/s]

30 h100
20 h500
10
0
5 3 1 1 3 5
y/h [ ]
(b)

0.5
0.4
[]
O ,RMS

0.3 h100
0.2 h500
2
Y

0.1
0
5 3 1 1 3 5
y/h [ ]
(c)

Figure 3.27: RMS profiles, at x = 5h for the h100 and h500 meshes, of (a) axial velocity,
(b) transverse velocity and (c) O2 mass fraction.
3.3. Reacting Flow 63

3.3 Reacting Flow


In this section, the reacting flow field is studied. The cold flow solution is instantly
replaced by a hot flow field following the procedure detailed in Sec. 3.1.6. While keeping
the main characteristics of the non-reacting flow in mind, the following questions are
addressed:

Where is the flame stabilized and how ?

Although the reactants are non-premixed, is the turbulent flame a pure diusion
flame ? How can this flame structure be modeled in a LES computation ?

Are the comb-like structures still forming in the reacting flow and if so, how do they
interact with the flame ?

3.3.1 Flame stabilization


An instantaneous temperature field is presented in Fig. 3.28 with superimposed iso-
contours of density gradient (green=4 107 kg/m4 ) and heat release rate (grey=1012 W/m3 ,
black=1013 W/m3 ). The flame is not lifted and the hot gases touch the splitter plate.
This is partially due to the adiabatic boundary condition imposed at the solid boundary.
However, H2 /O2 flames are known to be extremely reactive and experimental evidence
under comparable thermodynamic conditions showed that the flame is stabilized very
close to the splitter [Singla et al. 2007].

Figure 3.28: Temperature field with superimposed density gradient (green) and heat-
release (black: max heat release rate of case AVBP_RG (=1013 W/m3 ); grey: 10 % of
case AVBP_RG (=1012 W/m3 ).

The temperature field is much wrinkled on both oxygen and hydrogen sides. Pock-
ets of oxygen are sometimes captured in the hot gases, especially in the second half of
the computational domain. This observation allows to speculate that these pockets of
warmed-up oxidizer could eventually enter the hydrogen stream and generate a corre-
sponding unsteady heat release. The combustion modes will be studied in more details
in Sec. 3.3.3.
Chapter 3. A DNS study of turbulent mixing and combustion in the
64 near-injector region of Liquid Rocket Engines

The green iso-contour highlights the region of high density gradient and allows to
distinguish the dense stream from the lighter regions.
A close-up view of the vicinity of the splitter is presented in Fig. 3.29. Regions of
strong reaction rate are visible at the edges of the high-temperature pockets, both on the
hydrogen and oxygen sides, which shows the importance of turbulent transport across the
splitter height.

Figure 3.29: Close-up view of the flame stabilization zone behind the splitter. Temper-
ature field with superimposed iso-contours of density gradient (green=4 107 kg/m4 ) and
heat-release rate (grey=1012 W/m3 , black=1013 W/m3 ).

A schematic view of the flame splits it into two parts, as shown in Fig. 3.30:
a near-injector steady diusion flame between O2 and a mixture of combustion
products and fresh hydrogen (mainly H2 + H2 O), at the LOx corner. The flame is
quasi-laminar because it is not disturbed by the vortices shed at the H2 corner.

a turbulent diusion flame developing further downstream.

3.3.2 Vortex shedding and comb-like structures


Figure 3.31 shows the density field for both the non-reacting and the reacting flow. The
first striking feature is the dierence between the LOx stream shapes: the LOx stream
moves straight forward in the hot flow case whereas it is rolled up by the large vortices shed
at the H2 corner of the lip in the cold flow case. These complex flame/vortex interactions
drive turbulent mixing. Such reduction in the vertical extent of the mixing layer under
reacting conditions has already been observed at low pressure [Renard et al. 2000].

Curvature of the density field


In order to quantify the visual dierences observed in the reacting and non-reacting
cases, the probability density function (PDF) of the curvature of a density iso-surface is
presented in Fig. 3.32. The value chosen for the curvature analysis is the iso-surface at
3.3. Reacting Flow 65

Vortex shedding
recirculation zone
H2 (H2O + H2)

Steady Turbulent
O2 Flame Flame

Figure 3.30: The flame/vortex interaction separates the flame into 2 zones, a near-injector
steady diusion flame between O2 and a mixture of H2 and H2 O, and a turbulent flame
developing further downstream.

the intermediate density between the hydrogen and oxygen streams 0.5 ( 600 kg.m3 ),
defined in Eq. 3.15. While the PDF for the reacting flow is roughly symmetric and
centered around a zero value, the curvature PDF for the cold flow is double peaked: the
peak at 104 m1 corresponds to a concave radius of curvature of h/5 in the density
isocontour. This indicates that although visually, the large density waves appears to have
a characteristic radius of curvature of h, they actually consist in a collection of highly
curved pieces. The dierence between the hot and cold flow curvature PDF also confirms
the visual observation that the density waves disappear in the hot flow. Note that the
PDFs plotted in Fig. 3.32 are actually shown in the [-1 105 , 1 105 ] subrange, which allows
to clearly observe the double-peaked PDF in the cold case. The integrals of the PDFs
between and + are both equal to one.
Chapter 3. A DNS study of turbulent mixing and combustion in the
66 near-injector region of Liquid Rocket Engines

(a)

(b)

Figure 3.31: Comparison of the density fields between a) the non-reacting flow and b) the
reacting flow

Figure 3.32: Curvature PDF of the median density (0.5 iso-contour for the cold and
reacting flows.
3.3. Reacting Flow 67

3.3.3 Combustion Modes


Direct visualization
Figure 3.33 shows the fields of oxygen, hydrogen and water (major species), with super-
imposed stoichiometric mixture fraction isoline. The flame is predominantly composed of
non-premixed flame elements located on the stoichiometric mixture fraction line, to which
pure hydrogen flows against pure oxygen.

H2O is flushed by H2

H2 pockets trapped between


merging oxygen fingers

(a)

Formation of O2 pockets from fingers

Diluted O2 pocket that has crossed Zst

(b)

H2O is flushed by H2

Trapped pockets of
combustion products

(c)

Figure 3.33: Reacting flow: instantaneous fields of (a) hydrogen (b) oxygen and (c) water
mass fraction, with superimposed stoichiometric mixture fraction isoline (black). The
dash line in (b) shows the location of the 1D cut that is used in Fig. 3.34.

In a similar manner as in the non-reacting flow, the passage of the vortices shed at
the lip, rolls the oxygen stream and initiates the formation of oxygen fingers, as shown
Chapter 3. A DNS study of turbulent mixing and combustion in the
68 near-injector region of Liquid Rocket Engines

in Fig. 3.33(b). The exchange surface between oxygen and hydrogen is thus increased,
which enhances the consumption rate of reactants. The fingers of oxygen are sometimes
sliced into isolated pockets, due to the joint action of turbulence and chemical reactions.
A pocket of oxygen appears above the stoichiometric line on the left of Fig. 3.33(b) (dash
line), because of dilution with the surrounding combustion products, as shown in Fig. 3.34.
Pockets of hydrogen and pockets of combustion products can also be trapped into the
oxygen stream as shown in Fig. 3.33, for instance when two fingers of oxygen merge.
Immediately behind the lip, the fluid is composed of a large amount of recirculated
combustion products and a small amount of fresh hydrogen that is continuously supplied
by coherent structures. Intermittently, a large coherent structure can flush this pocket of
combustion products which is replaced by fresh hydrogen, as shown in Fig. 3.33(a). Thus,
the reservoir of fuel of the steady diusion flame that is attached to the splitter plate, has
a varying composition which induce an oscillatory flame structure.
As already observed in the non-reacting case, turbulence develops with the downstream
distance of the lip which makes flame/turbulence more important. Before the end of the
computational domain, this creates mixed combustion modes where pockets of hydrogen
(or oxygen) trapped into the oxygen stream (or the hydrogen stream) can burn against a
non-premixed flame element, which creates complex flame structures.

Flame structure in mixture fraction space


The reacting flow is now observed in the mixture fraction space, in Fig. 3.35, where the
counterflow flame computed in Sec. 3.1.7 is also shown as a reference (dashed red line).
Consistently with the direct visualization of the flame, no extinction occurs and the
great majority of points are located along the diusion flame manifold. The adiabatic
flame temperature for a stoichiometric mixture is approximately 3800 K and this value is
reached at the stoichiometric mixture fraction.
From visual observations of the unsteady flame (not shown), it has been found that
the points located above equilibrium are associated with highly curved pockets of reac-
tants. The preferential diusion eects, which are enhanced by curvature, can induce
super-equilibrium states as shown for instance in [Gicquel et al. 2005]. Since far from
extinction, the variation with strain rate of the maximum temperature in a high pres-
sure H2 /O2 counterflow diusion flame is small (Tmax < 100 K at 10 MPa, over 7
orders of magnitude of the strain rate [Ribert et al. 2008]), the points below the reference
temperature are induced by dilution.
It is therefore concluded that even though chemistry plays a key role in the stabilization
of such flame, turbulent transport is also active and influences the combustion modes. It
is also foreseen that for fuels with less chemical activity, turbulence could become a key
stabilization process.
3.3. Reacting Flow 69

1 H2
O
Y[] 2
0.5 H2O

0
0 0.2 0.4 0.6 0.8 1
x/h [ ]
(a)

0.2
O
0.15 H 10
OH
Y[]

0.1

0.05

0
0 0.2 0.4 0.6 0.8 1
x/h [ ]
(b)

4
4 x 10 H2O2
HO2
Y[]

0
0 0.2 0.4 0.6 0.8 1
x/h [ ]
(c)

Figure 3.34: Cut through a pocket of oxygen diluted with combustion products. The
location of the 1D cut is shown by a dash line in Fig. 3.33(b).
Chapter 3. A DNS study of turbulent mixing and combustion in the
70 near-injector region of Liquid Rocket Engines

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

Figure 3.35: Flame structure in the mixture fraction space. counterflow diusion
flame at a = 3800 s , scatter plot of the present turbulent flame.
1
3.3. Reacting Flow 71

3.3.4 Mean flow field


The reacting flow field was averaged over a period of 5 Tf t ( 0.6 ms), which is large
enough to observe the main characteristics of the mean field even though the statistics
may not be fully converged. Figure 3.36 shows the resulting average density field, using the
same logarithmic scale and iso-contour as for Fig. 3.20. The outlet region corresponding
to the sponge-layer described in Sec. 3.1 is grayed out because the flame characteristics
are not physical in this region. For the present reacting flow, the angle of the dense core
is 09 = 2 , which corresponds to a dense core length L 30 R. This value, based on
a relatively short computational domain, is three times larger than the cold flow. This
is attributed to the heat release rate, which prevents the formation of the large vortices
observed in the cold flow (c.f. Fig. 3.28).

Figure 3.36: Average density field for the reacting flow. White iso-contour: = 0.9 in
O2 .

Figure 3.37(a) and Fig. 3.37(b) show the mean temperature and heat release rate
fields, respectively. While the temperature is relatively homogeneous behind the splitter
plate (because of turbulent mixing), the mean reaction rate clearly highlights two regions,
as already depicted in Fig. 3.30: (1) a relatively steady flame is anchored at the lower
corner of the lip (oxygen stream) and (2) an unsteady flame brush, with a vertical extent
comparable to that of the splitter, is observed further downstream. It seems that in region
(1), chemistry strongly dominates while in region (2), the turbulence level is high enough
to control the flame front.
Chapter 3. A DNS study of turbulent mixing and combustion in the
72 near-injector region of Liquid Rocket Engines

(a)

(b)

Figure 3.37: Average fields for the reacting flow. (a) Temperature. (b) Heat release rate.
Black iso-contour: 0.9 .
3.3. Reacting Flow 73

3.3.5 Comparison of numerical results with existing experimental


data
Observing the flame dynamics in a high-pressure environment is a dicult task and to the
authors knowledge, only the research teams of the EM2C laboratory have been able to
produce detailed measurements of both OH and OH Planar Laser Induced Fluorescence
(PLIF) for transcritical injections [Juniper et al. 2003,Singla et al. 2005,Candel et al. 2006,
Singla et al. 2007].
In an attempt to compare the present numerical results with experimental flame vi-
sualizations, the OH mass-fraction fields at four instants is plotted in Fig. 3.38 and the
OH-PLIF images of a transcritical H2 /O2 flame at 6 MPa from [Singla et al. 2007] are
shown in Fig. 3.39. In Fig. 3.38, the various combustion modes identified in Sec. 3.3.3 are
clearly observed: strained-diusion-flame elements and pocket combustion in the near-
field region, and mixed combustion modes further downstream. In the PLIF images of
Fig. 3.39, the flame appears much thicker and no such complex flame dynamics is ob-
served. The numerical flame thickness is consistent with 1D laminar counterflow diusion
flames at strain rates comparable to the velocity gradient at injection that have a thick-
ness ranging from = 80 m to = 30 m between a = 19 103 s1 and a = 190 103 s1 .
Similarly, the numerical study of [Ribert et al. 2008] show that the flame thickness is
small, even at low strain values (for instance, = 200 m, with a = 2000 s1 , at 10 MPa)
which is in the line of the present numerical results and tends to indicate that the flame
should not be as thick as the splitter plate.

H2

O2

(a) (b)

(c) (d)

Figure 3.38: OH mass-fraction fields at four instants. Liquid oxygen and gaseous hydrogen
are injected below and above the step, respectively.

To understand the discrepancy between simulation and experiment, it is then necessary


to go back to the measurement technique. In a high-pressure reacting environment, due to
the high Reynolds number, turbulent scales are smaller than at atmospheric pressure, and
Chapter 3. A DNS study of turbulent mixing and combustion in the
74 near-injector region of Liquid Rocket Engines

O2

H2

Figure 3.39: OH-PLIF images of the flame holding region. Liquid oxygen and gaseous
hydrogen are injected above and below the step, respectively (opposite arrangement of
Fig. 3.38). The OH distribution in the flame edge is shown on the standard color scale.
p = 6.3 MPa (fig. 5 of [Singla et al. 2007])

require an increased level of spatial resolution and data acquisition rate. Unfortunately,
optical diagnostics face several hurdles at high-pressures. The first one is light scattering
due to large density gradients, which can deviate light path and thus optical signal. The
second one is collisional quenching, which degrades the quality of PLIF. Due to the high
density of cryogenic fluids, the excited state of radical species induced by laser sheet can
relax to its equilibrium state without fluorescence emission, due to kinetic energy exchange
while molecules are colliding together. This results in a low signal-to-noise ratio.
PLIF is also subject to blurring eects caused by curvature of the turbulent structures
intercepted by the laser sheet, which is not infinitely thin. In [Singla et al. 2007], the
thickness of the laser beam is 0.5 mm, which is equal to the splitter plate height in the
present configuration.
Thus, to compare PLIF results with numerical results, the heat release rate should be
integrated over a 1h distance in the spanwise direction. As the present computation is
2D, instead of a line-of-sight spatial average, a time average is shown in Fig. 3.40. The
thickness of the time-averaged flame is now comparable to the splitter plate height, as in
the experiment. This suggests that the PLIF measurements do not show a visualization
of a cut through the flame front, but rather the flame brush due to spatial averaging in
the spanwise direction. The thin diusion flame located immediately behind the splitter
plate, is very thin on the numerical results and may be within the noise of the PLIF
signal. This could explain why the experiment seems to show a lifted flame, while the
simulation shows an attached flame. However, further experimental studies would be
needed to substantiate this claim.
3.4. Conclusions 75

Figure 3.40: Comparison between experimental signal of OH PLIF [Singla et al. 2007] at
transcritical conditions and time averaged heat release rate in the current simulation (the
numerical visualization is upside down to compare with experimental visualization)

3.4 Conclusions
Simulations of a splitter-plate configuration representative of an injector of a Liquid
Rocket Engine have been conducted. The dynamic and thermodynamic conditions are
chosen to be representative of a real engine under steady operation. Both non-reacting and
reacting configurations were computed, with a two-dimensional domain and full resolution
of turbulence and chemistry.
Vortex shedding behind the injector lip has been shown to play an important role in
the mixing of reactants, and is responsible for the formation of finger-like structures at
the surface of the oxygen jet, which exhibits both large and small-scale structures. The
Strouhal number of vortex shedding is St = 0.2, based on the convection speed of the
coherent structures and their wavelength. In the non-reacting case, 100 grid points are
sucient to discretize the flow in the wake of the splitter plate and to obtain converged
statistics.
In the reacting case, the large scale structures disappear in the presence of hot combus-
tion products between the oxygen and hydrogen streams, which results in a less intense
turbulent mixing that lowers the spreading angle of the oxygen jet. The flame is anchored
at the splitter plate and is separated into two main regions: a steady diusion flame be-
tween pure oxygen and a mixture of hydrogen diluted with combustion products, which
sticks to the splitter plate, and a turbulent diusion flame further downstream, containing
strained non-premixed flame elements and burning pockets of reactants. The Damkhler
number is suciently far from extinction so that the flame structure is similar to the
H2 /O2 non-premixed diusion flame manifold, with a small scatter induced by dilution
of the reservoirs by combustion products. This shows that one can legitimately use the
equilibrium assumption in a turbulent combustion model for such a flow configuration.
The comparison of the present flame visualizations and OH-PLIF images tends to
indicate that the flame thickness is actually much thiner than what is measured exper-
imentally. This might stems from the finite-thickness of the laser sheet that induces a
Chapter 3. A DNS study of turbulent mixing and combustion in the
76 near-injector region of Liquid Rocket Engines

spatial integration of the reaction zone in the spanwise direction.


Chapter 4
Large Eddy Simulations of a
Transcritical H2/O2 Jet Flame Issuing
from a Coaxial Injector, with and
without Inner Recess

4.1 Introduction
Most high-performance Liquid Rocket Engines (LRE) are equipped with coaxial injectors,
where a central dense jet is atomized by a high-speed outer jet, due to aerodynamic forces.
The operating principle of these coaxial injectors is the same as air-blast atomizers, which
are classical in aeronautical injection systems. As indicated by [Candel et al. 2006], sys-
tematic tests carried out on a wide variety of injection devices led to rapid adoption of
such configurations, which ensure a smooth ignition transient and a stable nominal opera-
tion. In the past 50 years, most design parameters have been adjusted without a detailed
understanding of flame dynamics, because of both limited experimental diagnostics and
numerical capabilities.
About twenty years ago, a research-test facility for high pressure combustion of cryo-
genic propellants was built at ONERA (France) and incrementally upgraded over the
years. Many experimental studies have been carried out in this facility, named Mas-
cotte [Habiballah et al. 1996,Candel et al. 1998,Vingert et al. 1998,Habiballah et al. 1998,
Haberzettl et al. 2000, Vingert et al. 2000, Pourouchottamane et al. 2001, CNES 2001,
Vingert et al. 2002]. See [Habiballah et al. 2006] for a detailed description of the test
bench, in its current version.
Taking advantage of the optical access to the combustion chamber, advanced laser
techniques (mainly Coherent Anti-stokes Raman Spectroscopy (CARS) and Planar Laser
Induced Fluorescence (PLIF)) and line-of-sight imaging (shadowgraphs and OH* emis-
sion) have allowed a precise monitoring of LOx/GH2 and LOx/LCH4 flame shapes and
lengths, as well as an identification of controlling factors. Table 3 in [Candel et al. 2006]
gives an overview of imaging studies conducted at Mascotte.
At supercritical pressure, the break-up of the LOx jet was shown to follow the fibrous
regime, depicted in [Chigier & Farago 1992], where disturbances with small wavelengths
develop at the LOx surface, due to the peeling induced by the high-speed outer jet. Two
key parameters were identified: the momentum-flux ratio J = H2 u2H2 /O2 u2O2 and the
recess length of the inner oxygen tube. Both an increase in J and in the inner recess
length provokes a thicker flame brush and a shorter flame.
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
78 Issuing from a Coaxial Injector, with and without Inner Recess

Above a critical value of the momentum-flux ratio, located around 10, combustion
eciency is greatly enhanced [Herding 1997].
The eects of the recess length have been studied in [Kendrick et al. 1999]. It was
found experimentally that for operating points with a low momentum-flux ratio (the so-
called C operating points), recessing the inner LOx tube induced a better combustion
eciency, whereas such enhancement was not clearly observed on operating points with
a high momentum-flux ratio (A operating points). A theoretical explanation has been
proposed: the presence of the flame inside the recessed injector creates a reduction in the
flow section of hydrogen, which results in an acceleration, increasing the momentum-flux
ratio. This, in turn, greatly improves combustion eciency if the J = 10 critical value is
attained.
Another experimental and theoretical study from [Juniper et al. 2001,Juniper & Can-
del 2003b], has shown that even in the cold flow case, recessing the inner injection tube
can trigger a wake-like absolute instability, which enhances far-field mixing and creates
large sinuous wavelengths at the surface of the LOx surface.
On the numerical side, most studies have been conducted with RANS, to validate
models against experimental results [Cheng & Farmer 2006, Poschner & Pfitzner 2009,
Demoulin et al. 2009, Cutrone et al. 2010, Matuszewski 2011, Pohl et al. 2011]. Validating
RANS models in the real-gas framework is crucial, since validated models can then be used
for parametric design optimization in LRE development cycles. However, RANS is not
able to capture unsteady features of the flow field which can have a great impact on the
prediction of combustion eciency, such as the wake-like absolute instability mentioned
above, which justifies the use of LES in subsequent sections.
Only a few numerical studies have used LES as a tool for detailed understanding of
flame dynamics inside LREs. For instance, [Masquelet et al. 2009] led a 2D-axisymmetric
simulation of a multiple-injector combustor, and focused on wall heat flux induced by
combustion. Transcritical LES of coaxial injectors have been conducted in [Schmitt
et al. 2009, Schmitt et al. 2010a] and [Matsuyama et al. 2006, Matsuyama et al. 2010],
where the flow complexity has been captured with great precision and good agreement
with numerical experiments. These studies have shown the potential of LES in the context
of LRE studies.
In [Schmitt 2009], the eects of the momentum-flux ratio increase between the A60
and C60 operating points were studied with LES, using the AVBP code, also used herein.
The increased combustion eciency in the A60 case was mainly attributed to an increased
level of turbulent velocity fluctuations throughout the flame length, induced by the higher
momentum-flux ratio. This shortens the dense core length by 20% as well as the flame
length by 50%. Thus it has been found that both the near-field mixing and the far-field
mixing are influenced by the momentum-flux ratio.
In the present chapter, the eects of the inner recess length on combustion charac-
teristics is studied with two LES simulations, one without recess, and one with a recess
length equal to one diameter of the LOx tube dO2 , at the C60 operating point.
The objective is first to verify if, in agreement with theoretical derivations, an increase
of momentum-flux ratio is indeed responsible for mixing enhancements. Then, a detailed
description of the reacting flow field will complement the available experimental results
4.2. Configuration and operating point 79

and provide more insight into turbulent combustion processes at stake inside a LRE.

4.2 Configuration and operating point


4.2.1 The Mascotte facility
In its current version, the Mascotte test bench enables the study of LOx/GH2 , LOx/LH2
and LOx/LCH4 combustion at pressure ranging from 0.1 to 7 M P a. A square combustion
chamber, with movable lateral windows, can be used for optical access to the combustion
zone, as shown in Fig. 4.1. The combustion chamber is 400 mm long and the square
section has a 50 mm side length. A coaxial injector is fed-in with oxygen at a maximum
mass flow rate of 450 g/s, and a maximum hydrogen mass flow rate of 75 g/s. At the
outlet, a chocked nozzle enables the adjustment of the chamber pressure. In the numerical
simulations presented hereafter, this part is not included and pressure is imposed before
the converging part of the nozzle.

400 mm

Coaxial 50 mm
injector

Movable lateral windows


Jet Flame

Figure 4.1: Square combustion chamber with optical access used in the Mascotte facility.

4.2.2 C60 operating point


For the C60 operating point, the combustion chamber pressure is 6 MPa, and the mo-
mentum flux ratio is below the J = 10 critical value. More details about the injection
conditions is given in Sec. 4.3.2. The low momentum flux value of the C60 operating
point makes the recess length eects more important, which is suitable to understand
the basis mechanisms for combustion enhancement. The thermal power delivered by
combustion at the C60 operating point is 1.5 MW (= mO2 1+s s
h0f,H2 O ), which is represen-
tative of the power delivered by one of the 566 coaxial injectors of the Vulcain 2 engine
(= 1.5 109 /566 = 2.6 MW). Thus, understanding how the recess enhances combustion for
this lab-scale flame might provide useful guidelines for the design of LRE injectors.

4.2.3 Characteristic numbers and scales


Figure 4.2(a) shows the dimensions of the non-recessed coaxial injector. The lip height has
been increased from 0.3 mm in the experiments to 1.0 mm in the numerical simulations,
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
80 Issuing from a Coaxial Injector, with and without Inner Recess

to allow a sucient discretization of this flame stabilization zone and prevent numerical
stiness. If the oxygen flow is attached to the diverging part of the injector in the ex-
periments, then the oxygen velocity would decrease and the momentum-flux ratio at the
injection plane would be twice larger in the experiment than in the numerical simulation,
although still smaller than the J = 10 critical value (see Tab. 4.1). This geometrical mod-
ification could thus have an impact on the near-injector flame dynamics, but still allows
the study of the eects of recess. Moreover, a similar geometrical modification was done
in [Schmitt et al. 2009] and have yielded good qualitative agreement with experimental
results. The inner recess-length is shown on Fig. 4.2(b). It is the distance between the
exit of the inner oxygen tube and the exit of the outer hydrogen tube. In what follows,
NR and R will refer to the non-recessed and the recessed configurations, respectively.

(a) (b)

Figure 4.2: Geometry of the coaxial injectors: a) dimensions and b) recess length.

Using Eq. 3.5, the convection speed of coherent structures in the inner shear layer is
Uconv = 23 m/s. The characteristic convection time is: conv = h/Uconv = 43 s.

4.3 Numerical setup


Mesh
Two dierent mesh resolutions have been used:
a first mesh, called h5, with 5 cells within the lip height h and a total of 2 million
nodes and 12 million tetrahedra, is used to compute the mean and RMS values for
the whole flow field.

a second mesh, called h10, with 10 cells within the lip height and a total of 21
million nodes and 120 million tetrahedra has then been used to check wether the
global hydrodynamic instability, that is observed experimentally, is triggered with
a finer mesh.
Figure 4.3 shows a cut of the h5 mesh, which is essentially composed of two regions: a
near-injector region with a constant spacing = h/5 and a progressive mesh coarsening
4.3. Numerical setup 81

following the opening of the hydrogen jet. Most of the grid points are placed in the near-
injector region because the dierences between the NR and R configurations presumably
originate from this region. The h10 mesh is a homogeneous refinement of the h5 mesh,
with characteristic lengths divided by 2 in each direction.

Figure 4.3: Cut of the h5 mesh, colored by = V 1/3 , with V the local volume of cells.

Most numerical results shown in the present chapter uses the h10 mesh. When the h5
mesh is used, it is explicitly noted in the text and the legends of figures.

CPU cost
Using the flame length L = 70 dO2 in Eq. 2.49, the hydrogen flow-through time is
Tf t,H2 = 10 ms and the oxygen flow through time is Tf t,H2 = 250 ms.
With the h5 mesh, the time step is dt = 3.3 108 s and the number of iterations needed
to simulate 30 Tf t,H2 ( 1 Tf t,O2 ) is N30,f t,H2 = 106 (see Eq. 2.50). Thus, the simulation
cost is approximately 50 103 CPU hours (see Eq. 2.51), using the eciency of AVBP on
the VARGAS supercomputer at Idris (Power 6 processors at 4.7 GHz) Ci,n = 70 s.
With the h10 mesh, the cost is 24 times greater and requires 800 103 CPU hours to
obtain converged statistics throughout the flame length. At the moment, only = 2 ms
have been simulated. Thus, the length on which statistics are converged is smaller than
70 dO2 :

uH 2
L = (4.1)
10
= 15 dO2

and only the statistics of the near-injector region will be analyzed with the h10 mesh.
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
82 Issuing from a Coaxial Injector, with and without Inner Recess

Numerical scheme
A two-step Taylor-Galerkin finite-element scheme is used, called TTG4A [Colin & Rudg-
yard 2000]. It is third order in space and time.

4.3.1 Models
Equation of State (EOS)
The Soave-Redlich-Kwong (SRK) EOS is chosen here, because it allows a better rep-
resentation of densities at low temperatures (far below the pseudo-boiling temperature)
than the Peng-Robinson (PR) EOS (see Fig. 2.3). Thus, both the mass flow rate and the
velocity of a transcritical injection are well represented using the SRK EOS. This would
not be the case with the PR EOS as, due to the over-estimation of density, the injection
velocity would be lowered to maintain the mass flow rate, possibly modifying the dynamic
behavior of the coaxial jets.

Molecular Diusion
As will be seen later on, the integral length scale is of the order of lt = h and the
velocity fluctuations are of the order of u = 70 m/s (= 0.3 uH2 ) (see Fig. 4.12). Using
the kinematic viscosity of the hydrogen stream and assuming a fully developed turbulent
flow, the Kolmogorov scale is = 0.3 106 m (see Eq. 3.7).
On the other hand, the mesh filter size is = 100 m ( = 200 m) for the h10 (h5)
mesh. Thus , so that molecular diusion is neglected in this chapter.

Turbulent Combustion Model


The turbulent combustion model that is used herein has been developed and validated
in [Schmitt et al. 2010a], where both an atmospheric H2 /air and a transcritical LOx/GCH4
jet flames have been convincingly compared to experimental data.
This model is based on the presumed beta function subgrid LES approach, including
heat release in the fast chemistry limit [Peters 2001].
In Chap. 3, it has been shown that, at the stabilization point of a transcritical
LOx/GH2 flame, the scalar dissipation rate is suciently far from the extinction value,
so that the flame structure is close to the non-premixed diusion flame manifold and the
equilibrium lines (see Figs. 3.35 and 3.7). Downstream of the stabilization point, scalar
dissipation rates are lowered by turbulent mixing, as exemplified in Fig. 3.18. Thus,
modeling combustion in the fast chemistry limit appears to be justified.
A single step reaction is assumed:
mH2 + s mO2 (1 + s)mH2 O (4.2)
with mk the mass of the k reactant and s the stoichiometric mass ratio (=8 for H2 /O2
combustion).
For H2 /O2 combustion considered herein, 3 species are transported: H2 , O2 and H2 O.
The mixture fraction is computed from the mass fractions of these species.
sYH2 YO2 + YO02
Z= , (4.3)
sYH02 + YO02
4.3. Numerical setup 83

where the superscript 0 refers to the injection conditions.


For laminar flows, the Burke and Schumann solution [Burke & Schumann 1928] is
directly used, where species mass fractions are piecewise-linear functions of the mixture
fraction, as shown in Fig. 4.4. A source term is then computed for combustion products
to relax the transported species mass fractions Yk towards the equilibrium lines Yke :
Yk Yke
k = (4.4)

where is a relaxation time (a few acoustic time steps).

1
H
2
Y[]

O
0.5 2
H2O

0
0 zst
0.2 0.4 0.6 0.8 1
Mixture Fraction
Figure 4.4: Species mass fraction as a function of mixture fraction, in the infinitely fast
chemistry limit (Burke-Schumann solution).

For filtered turbulent flows, subgrid-scale flame/turbulence interactions are modeled


by convolution of the Burke and Schmumann solution with the presumed-shape beta PDF
P , that depends on the filtered mixture fraction Z and on the subgrid-scale variance Z :
1
e
Yk (Z, Z ) = Yk (Z )PZ,
Z

(Z )dZ

(4.5)
0

The values of Y e
k (Z, Z ) are stored in a precomputed table that is accessed during the
LES computation. The model requires to solve the transport equation for Z, while an
[Pierce & Moin 1998, Peters 2001]:
algebraic expression is used to determine Z
Z 2
= C2 |grad(Z)| (4.6)
with C a model constant. In [Schmitt et al. 2010a], the influence of this constant was
shown to be small within the 0.1-0.5 range. Here, C is set to 0.05. Finally, the filtered
species source terms are:
Yk Ye
k = k
(4.7)

and the filtered heat-release rate is:
N

T = h0f,k k (4.8)
k=1
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
84 Issuing from a Coaxial Injector, with and without Inner Recess

Note that this model is able to reproduce fast-burning stabilized flames only. Induction
zones or dilution phenomena, as seen in Chap. 3, can not be reproduced.

Turbulent Viscosity
The subgrid-scale turbulence model is the WALE model, which is described in Sec. 2.3.2.
This model sets turbulent viscosity to zero in pure-shear regions, which allows the devel-
opment of coherent structures by shear instability and goes to 0 near walls. This is
particularly important to capture the laminar to turbulent transition of jets. Turbulent
viscosity is activated in regions with either large strain rates, large rotation rates, or both.
This model has been used in LES of supercritical and transcritical non-reacting nitro-
gen jets in [Schmitt et al. 2010b], as well as transcritical LOx/GCH4 reacting flow [Schmitt
et al. 2010a], in which numerical results have been favorably compared to experimental
results.

4.3.2 Boundary Conditions


The pressure is imposed at 6 MPa, using NSCBC, before the converging nozzle, at the
outlet of the combustion chamber shown in Fig. 4.1. The main scales and non-dimensional
numbers characterizing the C60 operating point are summarized in Tab. 4.1. where R

T [K] m [g/s] inj [kg/m3 ] Uinj [m/s] [Pa.s] R R Re Ma


O2 83 100 1190 10 0.2 103 0.5 1.2 1.8 105 0.01
H2 275 45 5.1 234 7.4 106 8.3 4.7 3.6 105 0.18
R Ru J E
230 23 2.3 2.2
Uconv [m/s] conv [s] Tf t,O2 [ms] Tf t,H2 [ms]
23 43 25 1.1

Table 4.1: Characteristic flow quantities for the C60 operating point

and R (defined in Eqs. 1.9 and 1.10) are the reduced temperature and pressure, respec-
tively. The oxygen flow-through time Tf t,O2 and the hydrogen flow-through time Tf t,H2 ,
are computed with L = 70 dO2 and U = U inj in Eq. 2.50.
Oxygen is injected under transcritical conditions (R < 1, R >1), whereas hydrogen
is supercritical (R > 1, R > 1). This results in a large density ratio R of 80 (defined
in Eq. 1.14).
The velocity ratio Ru (defined in Eq. 1.16) is large and induces fast-growing shear
instabilities. The Mach number is very low in the oxygen stream and rather small in the
GH2 stream, so that compressibility eects are small.
The mixture ratio E, defined in Eq. 1.6, is lower than the stoichiometric mass ra-
tio (s=8).
4.3. Numerical setup 85

Inlet velocity fluctuations


Since the Reynolds number is very large, the flow inside the feeding lines is turbu-
lent. The Kraichnan/Celik synthetic turbulence method is used to inject velocity fluctu-
ations at the inlets of the simulation. This procedure was initially developed by [Kraich-
nan 1970, Smirnov et al. 2001] and has been adapted to compressible flows in [Guezennec
& Poinsot 2009]. The mean and RMS velocity profiles imposed at the inlets have been
determined by fitting 5th order polynomials to the results of a companion LES of a fully-
developed turbulent pipe flow at Re = 105 . The normalized velocity profiles are plotted
in Fig. 4.5. The polynomials for the mean axial velocity Cu and the RMS velocity fluctu-
ations in cylindrical coordinates CuRM S , Cur,RM S , Cu,RM S read:
Cu (yr ) = 30.21 yr6 + 70.68 yr5 62.39 yr4 + 25.48 yr3 4.921 yr2 + 0.3395 yr + 0.9951
CuRM S (yr ) = 289.9 yr6 + 724.1 yr5 678.8 yr4 + 296.8 yr3 57.82 yr2 + 4.657 yr + 0.9219
Cur,RM S (yr ) = 14.69 yr6 + 47.1 yr5 60.91 yr4 + 34.45 yr3 7.704 yr2 + 0.7724 yr + 0.9843
Cu,RM S (yr ) = 70.71 yr6 + 185.1 yr5 187.1 yr4 + 88.25 yr3 18.07 yr2 + 1.648 yr + 0.9685
(4.9)
where yr = (r rc )/R is a normalized distance from the injector wall, r is the distance
from the injector axis and R is the feeding line half-height. In the oxygen stream, rc = 0
and R = dO2 /2. In the hydrogen stream, rc = (ro,H2 + ri,H2 )/2, R = (ro,H2 ri,H2 )/2, ri,H2
and ro,H2 are the inner and outer radius of the H2 feeding line.

3
1
ui,RMS/ui,RMS

2 axial
c
u/uinj

radial
0.5
1 azimuthal

0 0
0 0.5 1 0 0.5 1
r/R [ ] r/R [ ]
(a) (b)

Figure 4.5: Turbulence injection profiles: a) mean axial velocity and b) Root Mean Square
velocity fluctuations. The superscript c refers to the centerline value.

The centerline mean axial velocity uinj is given in Tab. 4.1 and is determined from
the inlet mass flow rate. The centerline axial RMS velocities ucRM S are determined using
a 4% turbulent intensity, which is the value obtained in the companion pipe flow LES:
ucRM S
Itc
= inj (4.10)
u
Finally, the ratio between axial and radial or azimuthal centerline velocity fluctuations is
set to 0.65, which is the value obtained in the turbulent pipe flow LES.
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
86 Issuing from a Coaxial Injector, with and without Inner Recess

Initial Solution
The combustion chamber is initially filled with H2 O at rest and at T = Tadiab (Tadiab =
2200 K for the mixture ratio indicated in Tab. 4.1) and P = 6 MPa. Then H2 and O2 are
progressively injected into the chamber. The reactants mix and burn when they reach the
lip of the injector. The simulation is first run during 1 Tf t,O2 , which is the time needed
for the oxygen jet to settle in the combustion chamber, and allows the domain-averaged
quantities (mass fluxes, pressure, resolved kinetic energy, etc.) to reach a steady-state.
The averaging of the flow is then started.
This procedure is only used for the h5 mesh. An established flow field from the h5
mesh is interpolated on the h10 mesh, and is used as an initial solution. The averaging
of the flow is started directly after interpolation, since the flow field adapts rapidly to the
increased resolution.
4.4. Results 87

4.4 Results
First, the reference solution without recess is studied and validated against experimental
results. The recessed configuration is then studied.

4.4.1 Reference solution without recess


First, instantaneous results are shown, to qualitatively describe the flow features. Then,
time-averaged quantities are compared to experimental data.

4.4.1.1 Instantaneous Results


Fig. 4.6 shows cuts of instantaneous temperature and velocity fields. The flame is located
between the high-speed outer hydrogen jet and the low speed central oxygen jet. There
are essentially two regions in the flow field:

1. the near-injector region, in which the flame expands and pushes the hydrogen jet
away from the axis of the oxygen jet. Vortical structures grow while convected
downstream, as the hydrogen jet evolves towards fully-developed turbulence.

2. the far-field region, where vortical structures growth is prevented by the presence
of the chamber walls and where the flame terminates.

1 2

(a)

1 2

(b)

Figure 4.6: Instantaneous visualization of the reacting flow: a) temperature and b) axial
velocity fields.
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
88 Issuing from a Coaxial Injector, with and without Inner Recess

Due to the limited extent of the observation windows on the sides of the combustion
chamber (see Fig. 4.1), only one of the two regions could be observed for a given hot fire
test. Since in the second region, the flame expansion is hampered by the chamber walls,
most of the imaging studies have been conducted with the observation windows mounted
in front of the first region, where the flame expansion is driven by the injection conditions.
Thus, in subsection sections, we will essentially focus on this first region.
A first qualitative comparison between the instantaneous numerical results and the
experimental data presented in [Juniper 2001] is shown in Fig. 4.7. An OH field is
compared to an iso-surface of temperature (T=1500 K) since they both represent the
flame location. The flame is largely wrinkled in Fig. 4.7(b) and the visible wavelengths
of the perturbations at the flame surface increase with axial distance, as the turbulent
structures grow. The flame expansion appears to be retrieved in the numerical results.
A shadowgraph is compared to the mixture fraction field since both of them locate the
edges of the jets. The jets bear the same characteristics in the numerical simulation and in
the experiment: the hydrogen jet rapidly transitions towards turbulence while the oxygen
jet is more steady and largely penetrates the combustion chamber. Disturbances at the
surface of the oxygen jet develop because of the turbulent motions.
Figure 4.8(a) shows a cut of an instantaneous velocity field in the near-injector re-
gion. Coherent structures develop in the inner and outer shear layers, due to the Kelvin-
Helmholtz instability, which is qualitatively similar to what has been observed in the
DNS studies of Chap. 3 (see Fig. 3.8). The two shear layers then merge and the hydrogen
jet rapidly transitions towards developed turbulence, while the dense oxygen jet is very
stable and is convected along the axis of the injector. The flame is located in the inner
shear layer and is wrinkled by coherent structures and turbulence, as shown by a cut of
an instantaneous temperature field in Fig. 4.8(b).
Figure 4.9 shows a comparison between instantaneous experimental and numerical re-
sults, showing the exit of the injector. Although a good qualitative agreement is obtained,
the visual transition length D, needed for the vortices to develop after the injection plane,
seams slightly over-predicted.
4.4. Results 89

(a)

(b)

(c)

(d)

Figure 4.7: Qualitative comparison between instantaneous experimental ((a) and (c))
and numerical ((b) and (d)) visualizations of the reacting flow: (a) OH emission (b)
T = 1500 K iso-surface (c) shadowgraph (d) mixture fraction between 0 (black) and 1
(white).
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
90 Issuing from a Coaxial Injector, with and without Inner Recess

Developed turbulence

Outer shear layer

Inner shear layer

Coherent
Structures

(a)

(b)

Figure 4.8: Instantaneous a) axial velocity and b) temperature cuts.


4.4. Results 91

D D

(a) (b)

D D

(c) (d)

Figure 4.9: Qualitative comparison between instantaneous experimental ((a) and (c)) and
numerical ((b) and (d)) visualizations of the near-injector region of the reacting flow: (a)
OH emission (b) T = 1500 K iso-surface (c) shadowgraph (d) mixture fraction (black=0;
white=1).
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
92 Issuing from a Coaxial Injector, with and without Inner Recess

4.4.1.2 Time-averaged results


The objectives of this section are two fold. First, the comparison of time-averaged numer-
ical and experimental results serves as a validation. Then, the comparison of numerical
results obtained with the h5 and the h10 meshes allows to study the mesh convergence.
Figure 4.10 shows a cut of the average velocity magnitude (white=0 m/s; black=200 m/s)
with superimposed streamlines in black and density iso-contours in grey (=100, 500,
1000 kg/m3 ) obtained with the h5 mesh. A corner and a central recirculation zone are
generated by the impingement of the outer hydrogen jet on the chamber walls.

1000, 500, 100 kg/m3


Corner Recirculation Central Recirculation
Zone Zone

Figure 4.10: Cut of the average velocity magnitude (white = 0 m/s; black = 200 m/s)
with superimposed streamlines in black and density iso-contours in grey (=100, 500,
1000 kg/m3 ), obtained with the h5 mesh.

The fields of mean temperature are shown in Figs. 4.11(b) and 4.11(c), for the h5
and h10 meshes and are compared to the Abel-transformed OH emission of Fig. 4.11(a).
The overall flame shape is correctly captured by both meshes, which tends to indicate
that mesh convergence is already attained with the h5 mesh, for the NR configuration.
There are three distinct zones in the flame:

for x < 2dO2 , the flame is largely stretched by developing coherent structures in the
inner shear layer, and heat release is intense in this narrow region.

for 2dO2 < x < 20dO2 , a flame brush develops with turbulence.

for x = 20dO2 , because of the recirculation zones induced by the impingement of the
hydrogen jet on the chamber walls (see Fig. 4.10) the flame expansion increases.

Figure 4.12 shows a cut of the RMS fluctuations of the axial velocity. The initial growth
of the shear layers due to the formation of coherent structures is clearly observed, as well
as vortex breakdown (local peak of RMS velocities uRM S 70m/s at x 2dO2 ). In
the region of developed turbulence, the fluctuation levels slowly decrease as the turbulent
structures grow. It is noteworthy that turbulence mostly develops outside the dense
oxygen jet, in which the velocity fluctuations are very small. The density ratio between
the oxygen and hydrogen jets clearly limits the turbulent mixing of the two reactants,
4.4. Results 93

which is consistent with the DNS studies of Chap. 3 and with [Okongo & Bellan 2002b].
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
94 Issuing from a Coaxial Injector, with and without Inner Recess

(a)

(b)

(c)

Figure 4.11: Time-averaged flame shapes, shown by (a) Abel-transformed OH emission,


(b) mean temperature field with the h5 mesh, (c) mean temperature field with the h10
mesh (white=80 K; black=3200 K).
4.4. Results 95

Figure 4.12: Cut of the RMS fluctuations of the axial velocity.


Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
96 Issuing from a Coaxial Injector, with and without Inner Recess

CARS measurements
In [Habiballah et al. 2006] and [Zurbach 2006], temperature measurements have been
performed using Coherent Anti-Stokes Raman Spectroscopy (CARS), based on the spec-
tral signature of hydrogen. The validation rate of these CARS measurements depends on
the presence of hydrogen in the probe volume. Thus, the temperature measured by CARS
is not exactly the time-averaged temperature, but a conditional time-averaged tempera-
ture T , which can be defined as the mean temperature given that the mass fraction of
hydrogen YH2 is above a certain threshold :

T = T |YH2 > , (4.11)

where . is the time-average operator, and | is the conditional statement. Here, T is


obtained from 220 instantaneous temperature fields (with the h5 mesh), with varied
between 0 and 0.5, over a 12 ms period ( 10 Tf t,H2 ). The CARS mean temperature is
obtained from 225 samples over a run of 15 s. The results are plotted on Fig. 4.13. The
conditional mean temperature T is decreasing with , since hydrogen is in excess in the
combustion chamber and cools down combustion products.
From Fig. 4.13(a), it is clear that the probe lines at y = 4 mm and x = 50 mm cross
the flame front, whereas the corresponding temperatures obtained by CARS (the circles
in Figs. 4.13(b) and 4.13(d), respectively) do not seem to indicate the flame presence.
With = 0 (time-averaged temperature), a peak in the temperature profile is indeed
observed at the flame location. With = 0.3, the peak temperature is not observed
anymore because of hydrogen dilution, and a good agreement with experimental data is
obtained, which seems to indicate that the temperature measured by CARS is indeed
a conditional temperature. The agreement between the conditional mean temperatures
and the experimental data for the other probe lines, at x = 15 mm on Fig. 4.13(c) and
x = 100 mm on Fig. 4.13(e) is less conclusive, although the shapes of the T profiles
qualitatively follow the experimental trends.
4.4. Results 97

x = 15 mm x = 50 mm x = 100 mm

y = 4 mm

(a)

2500 _=0 y=4 mm 2500 x=15 mm


_=0

2000 2000
_=0
1500 1500
T [K]
T [K]

1000 _ = 0.3 1000

500 500
_ = 0.5 _ = 0.5
0 0
0 0.05 0.1 0.15 0.2 0.25 0 0.005 0.01 0.015 0.02 0.025
x [m] y [m]
(b) (c)

2500 x=50 mm x=100 mm


2500
_=0
2000 2000 _=0

1500 1500
T [K]

T [K]

1000 1000
_ = 0.3
500 500
_ = 0.5

0 0
0 0.005 0.01 0.015 0.02 0.025 0 0.005 0.01 0.015 0.02 0.025
y [m] y [m]
(d) (e)

Figure 4.13: NR configuration: comparison between H2 CARS measurements and condi-


tional temperatures T , defined in Eq. 4.11. : CARS measurements from [Habiballah
et al. 2006] and [Zurbach 2006]. (a) positions of the cuts (b) y = 4 mm, (c) x = 15 mm,
(d) x = 50 mm, (e) x = 100 mm.
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
98 Issuing from a Coaxial Injector, with and without Inner Recess

4.4.2 Eects of recess


Now that the flow field of the reference solution without recess has been characterized
and validated against experimental data, the recessed configuration can be analyzed.

4.4.2.1 Instantaneous Results


Figure 4.14 shows an instantaneous iso-surface of temperature (=1500 K) and a cut of
mixture fraction in the near-injector region of the R configuration. Coherent structures
develop first in the recessed part of the inner shear layer and disturb the outer shear layer
immediately at the injection plane. This provokes an early destabilization of the outer
shear layer and a greater opening of the hydrogen jet.

(a) (b)

Figure 4.14: Instantaneous visualization of the near-injector region in the R configuration:


(a) iso-surface of temperature (=1500 K) and (b) mixture fraction field.

An instantaneous experimental shadowgraph from [Candel et al. 2006] is compared to


the mixture fraction field in Figure 4.15. Although not as evidently as in the shadow-
graph, large wavelengths are observable at the surface of the oxygen jet, whereas they are
essentially absent from the NR configuration (see Fig. 4.7(d)).
4.4. Results 99

(a)

(b)

Figure 4.15: Qualitative comparison for the R configuration, between an instantaneous


(a) experimental shadowgraph and (b) a mixture fraction field between 0 (black) and 1
(white).
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
100 Issuing from a Coaxial Injector, with and without Inner Recess

4.4.2.2 Time-averaged Results


Figure 4.16 shows a comparison between time-averaged experimental and numerical re-
sults, in the near-injector region, for both the NR and R configurations. The recess length
eects on the flow field in the near-injector region are qualitatively captured. In both the
simulation and the experiment, the recess

decreases the visual potential core length D of the outer hydrogen jet and

increases slightly the opening angle of the outer hydrogen jet

The flame expansion is lower in the simulation than in the experiment. This might
indicate that the flame brush is not well predicted in the near-field region, or that the
mean temperature field do not directly correspond to the Abel-transformed OH emission.

To quantify the dierence in combustion eciency between the NR and R configura-


tion, the local oxygen consumption rate is integrated in a small control volume, positioned
at a varying distance x from the injector. The control volume is made of a slice perpen-
dicular to the axis of the injector, that is swept on a distance equal to 1 dO2 , as shown
in Fig. 4.17(a). The integrated consumption rate of oxygen O2 and the cumulated con-
sumption rate of oxygen I O2 read:

O2 (x) = O2 dV (4.12)
V (x)
x
I O2 (x) = O2 (x )dx , (4.13)
0

and are plotted on Figs. 4.17(b) and 4.17(c). In Fig. 4.17(c), the cumulated consumption
rate of oxygen is normalized by I O2 ,max = I O2 (xmax ), with xmax the maximum axial
distance from the injector, in the simulation. Between x/dO2 = 0 and x/dO2 = 2, the con-
sumption rate of oxygen increases very rapidly in both configurations, which corresponds
to the initial roll-up of the flame front by coherent structures. At x/dO2 = 0, a larger
amount of oxygen is consumed in the R configurations because coherent structures devel-
ops earlier in the R configuration than in the NR configuration, as shown in Figs. 4.9(d)
and 4.14(b), for instance. Between x/dO2 = 4 and x/dO2 = 8, the integrated consumption
rate of oxygen is locally larger in the NR configuration than in the R configuration, which
might come from a more intense vortex breakdown in the NR case. However, the oxygen
consumption rate is overall larger in the R configuration, as indicated on the cumulated
consumption rate of oxygen, on Fig. 4.17(c). Finally, the local consumption rate of oxy-
gen slowly vanishes at the axial positions further downstream, as oxygen is depleted. It
is clear from Fig. 4.17(c) that combustion eciency is globally increased by the recess of
the oxygen tube.
4.4. Results 101

D D

(a) (b)

D D

(c) (d)

Figure 4.16: Time-averaged results. NR configuration (left) and R configuration (right):


(a) and (b) are experimental shadowgraphs and Abel-transformed OH emissions, (c) and
(d) are mixture fraction fields (black=0; white=1) and temperature fields (black=1000 K;
yellow=3000 K).
Chapter 4. Large Eddy Simulations of a Transcritical H2 /O2 Jet Flame
102 Issuing from a Coaxial Injector, with and without Inner Recess

1 dO 2
(a)

0.01
NR
O [kg/s]

R
0.005
2

0
0 20x/d [ ] 40 60
O
2
(b)

1
,max
2

NR
IO / IO

0.5 R
2

0
0 20 x/d [ ] 40 60
O
2

(c)

Figure 4.17: (a) Definition and (b) distribution of the integrated consumption rate of
oxygen O2 . (c) Normalized and cumulated consumption rate of oxygen I O2 /I O2 ,max
(defined in Eq. 4.13)
4.5. Conclusions 103

4.4.2.3 Characteristic Numbers impacted by the recess


Following the theoretical derivations of [Kendrick et al. 1999], the increase of momentum-
flux ratio could provide an explanation for the enhancement of combustion eciency in
the R configuration. The local evolution of the momentum-flux ratio is thus plotted on
Fig. 4.18(a), and the local velocity ratio is plotted on Fig. 4.18(b). Local quantities
are obtained by taking the value at the center of each stream, at a given axial position.
Although the velocity ratio and the momentum-flux ratio are slightly increased at the
injection plane in the R configuration, the critical J = 10 value is not crossed and is thus
presumably not responsible for the combustion eciency enhancement in the simulations.

1.5
20
NR NR
1
R R
J[]

Ru [ ]
10
0.5

0 0
0 5 10 15 20 0 5 10 15 20
x/d []
O
2

(a) (b)

Figure 4.18: Axial evolution of (a) the momentum-flux ratio and (b) the velocity ratio.

4.5 Conclusions
Large Eddy Simulations have been conducted to study the eect of the recess of the inner
oxygen tube on the combustion eciency of a coaxial injector. First, the reference solution
without recess have been validated against experimental data and have allowed a precise
determination of the near-injector flame dynamics. Coherent structures develop in the
inner and outer shear layers due to the Kelvin-Helmholtz instability. The outer hydro-
gen jet rapidly transitions towards turbulence whereas the oxygen jet penetrates deeply
into the combustion chamber, because of the stabilizing eect of the large density ratio.
The post-processing of the numerical results also tends to indicate that the temperature
profiles measured by Coherent Anti-stokes Raman Spectroscopy (CARS) are conditional
temperatures rather than time-averaged temperatures.
The enhancement of combustion eciency have been retrieved in the numerical simu-
lations of the recessed configuration. The eects of recess on the momentum-flux ratio is
limited in the numerical simulations, and the most discernible eect is the shorter transi-
tion distance of the hydrogen jets, caused by the early development of coherent structures
in the inner shear layer. This favors downstream mixing of hydrogen and oxygen and
result in a better combustion eciency.
Chapter 5
General Conclusions

The objective of the present work was to study the characteristics of reacting flows at
operating conditions and in configurations relevant to Liquid Rocket Engines (LREs),
with high-fidelity unsteady numerical simulations.
The first hurdle to overcome is the transport of transcritical density gradients, which
requires numerical stabilization. This numerical filtering requires special attention in the
real-gas framework because of the non-linearity of the equation of state, which can induce
spurious velocity disturbances in the computed flow field. The solution that has originally
been proposed in [Schmitt 2009], is based on a consistent thermodynamic treatment of
the numerical dissipation terms and the theoretical background of this method has been
reinforced here.
In Chap. 3, a Direct Numerical Study of turbulent mixing and combustion in the
near-injector region of Liquid Rocket Engines has been conducted, at an unprecedented
resolution. Vortex shedding behind the injector lip has been shown to play an important
role in the mixing of reactants, and is responsible for the formation of finger-like structures
at the surface of the oxygen jet. The Strouhal number of vortex shedding is St = 0.2,
based on the convection speed of the coherent structures and their wavelength. This
non-dimensional frequency might play an important role, if coupled with the eigen modes
of the injection lines or the combustion chamber. In the reacting case, the flame stays
attached to the injector rim by a steady-diusion flame between the pure oxygen stream
and a mixture of hydrogen diluted by a varying level of recirculated hot gases. Further
downstream, the diusion flame is wrinkled by vortex shedding and the forming finger-like
structures are split into pockets of reactants that increase the reacting surface. Although
the combustion modes are visually complex, the global flame structure stay close to the
H2/O2 non-premixed manifold, with a small scatter induced by dilution of the unsteady
reservoirs with combustion products.
In Chap. 4, two Large Eddy Simulations of a transcritical H2 /O2 jet flame issuing from
a coaxial injector, with and without inner recess have been conducted. The combustion
enhancement induced by the inner recess of the oxygen tube has been retrieved in the sim-
ulations. The recess triggers the early development of coherent structures in the confined
inner shear layer that induces a better turbulent mixing further downstream. This shows
that LES is a promising tool for predicting in advance the eects of design parameters
and might play an important role in the development of next-generation injectors.
Appendix A
Thermodynamic derivatives

A.1 Getting the density from (P,T,Yi)


A.1.1 From the EOS to the cubic polynomial
Multiplying Eq. 2.46 by the two denominators of the RHS gives:

P (v b) D = RT D (v b)
P Dv D (P b + RT ) + v b = 0 (A.1)

Introducing en = (P b + RT ) and developing P Dv gives:



P Dv = P v 3 + d1 bv 2 + d2 b2 v
= P v 3 + P d1 bv 2 + P d2 b2 v (A.2)

Gathering the coecients gives:



P v 3 + (P d1 b en ) v 2 + P d2 b2 en d1 b + v en d2 b2 b = 0 (A.3)

The final cubic polynome reads:

a3 v 3 + a2 v 2 + a1 v + a0 = 0 (A.4)
with a0 = b ( + en d2 b)
a1 = P d 2 b2 e n d 1 b +
a2 = P d 1 b e n
a3 = P

NB : When using the perfect gas EOS, the coecients given in Eq. A.4 cannot be
used. Indeed, D = 1 in the perfect gas case, which totally changes the distribution of the
polynome coecients.

A.1.2 The Cardan method


The Cardan method gives the root of any third order polynome, so one can solve Eq. 2.47
directly to get the molar volume v. Although the solution of the Cardan Method is well
known (see for instance [Polyanin & Manzhirov 2007]), the demonstration is recalled here:
Start from a third order polynome with real coecients ai

a3 x3 + a2 x2 + a1 x + a0 = 0 (A.5)
108 Appendix A. Thermodynamic derivatives

Divide by a3 to obtain the normalized form


ai
x3 + b2 x2 + b1 x + b0 with bi = (A.6)
a3
Substitute x = y b2 /3 to eliminate the quadratic term

y 3 + py + q = 0 (A.7)
b22
with p = b1 (A.8)
3
3
b2 b1 b2
q = 2 + b0 (A.9)
3 3
Two new variables u and v are defined, so that:

y =u+v (A.10)

where v is NOT the molar volume here. As a new degree of freedom is introduced, a
relation between u and v must be introduced later. Inserting Eq. A.10 into Eq. A.9 gives:

u3 + 3uv (u + v) + v 3 + p (u + v) + q = 0
u3 + v 3 + (3uv + p) (u + v) + q = 0 (A.11)

Following Cardans idea, lets impose (3uv + p) = 0, in order to simplify Eq. A.11. This
condition has the following consequences:
p
(3uv + p) = 0 uv = (A.12)
3
p 3
u3 v 3 = (A.13)
3
Inserting Eq. A.12 into Eq. A.11 gives:

u3 + v 3 = q (A.14)

The sum and the product of u3 and v 3 are known using Eq. A.13 and Eq. A.14. Hence,
u3 and v 3 are the roots of the second order polynome
p 3
Z 2 + qZ = 0 (A.15)
3
(because (Z u3 )(Z v 3 ) = Z 2 (u3 + v 3 )Z + u3 v 3 ).
The discriminant of Eq. A.15 is
p 3
= q2 + 4 (A.16)
3
so that

q q
if 0 u = +3
and v =
3
(A.17)
2 2 2 2
q i q i
if < 0 u3 = + and v 3 = (A.18)
2 2 2 2
A.1. Getting the density from (P,T,Yi ) 109

As u and v are complex numbers, there are 3 solutions for u and v in Eq. A.17 and also
3 solutions for u and v in Eq. A.18.
1/3 2k
uk = u3 exp(i ) k = 0, 1, 2 (A.19)
3
For example, if > 0, Eq. A.17 implies
1/3
q 2 4
u1 = + or u2 = u1 exp(i ) or u3 = u1 exp(i ) (A.20)
2 2 3 3
1/3
q 2 4
v1 = or v2 = v1 exp(i ) or v3 = v1 exp(i ) (A.21)
2 2 3 3

However, two constraints lower the number of possible solutions for y:

y = u+v is a real number (A.22)


p
uv = (A.23)
3

If > 0
Then only the main roots (u1 , v1 ) (which are expressed in Eqs. A.20 and A.21) simul-
taneously satisfy the constraints of Eq. A.22 and Eq. A.23. Hence, if > 0 there is only
one real root:
b2
X1 = u 1 + v 1 (A.24)
3
1/3 1/3
q q b2
X1 = + + (A.25)
2 2 2 2 3

If < 0 3
First, lets notice that = q 2 + 4 p3 < 0 implies p < 0.
Then, lets express u3 in Eq. A.18 under exponential form:

u3 = U3 exp(i) (A.26)

q 2 p 3
3
U3 = ||u || = = (A.27)
2 4 3
q p 3
= arccos Re(u3 )/||u3 || = arccos / (A.28)
2 3

. So that the main cubic root of u3 reads:


3 1/3 1/3
u = U3 exp(i/3) (A.29)

p
= exp(i/3) (A.30)
3
110 Appendix A. Thermodynamic derivatives

Then Eq. A.19 can be written under exponential form:



p + 2k
uk = exp i
3 3
As y is a real number and is a sum of two complex numbers (y = u + v) then v is the
complex conjugate of u
y = u + u
. There are thus three real roots:
yk = uk + uk k = 0, 1, 2 (A.31)
= 2Re(uk ) k = 0, 1, 2 (A.32)

p + 2k
= 2 cos k = 0, 1, 2 (A.33)
3 3
Consequently, Eq. A.23 is satisfied
uk vk = uk uk = ||uk ||2 k = 0, 1, 2 (A.34)
p
= (A.35)
3
Finally, if < 0, the three real solutions of Eq. A.5 are:

p + 2k b2
Xk = 2 cos k = 0, 1, 2 (A.36)
3 3 3

q p 3
with = arccos / (A.37)
2 3

If = 0
Then the roots uk and vk are equal. The main roots read
q 1/3
u1 = v 1 = (A.38)
2
As = 0,
p 3
2
q +4 = 0
3
q 2 p 3
=
2 3
q 2/3 p
= (A.39)
2 3
. Thus q 2/3 p
2
u 1 v1 = u 1 = = (A.40)
2 3
and a first root is found:
y1 = u1 + v1 (A.41)
q 1/3
y1 = 2 (A.42)
2
A.1. Getting the density from (P,T,Yi ) 111

As u2 v3 = u3 v2 = u21 = p/3, a second real root exists:

y 2 = u 2 + v 3 = u 3 + v2 (A.43)
= u2 + u2 (A.44)
= 2Re(u2 ) (A.45)
= 2u1 cos (2/3) (A.46)
= u1 (A.47)

Finally, if = 0, the two solutions of Eq. A.5 are:


q 1/3 b
2
X1 = 2 (A.48)
2 3
q 1/3 b
2
X2 = (A.49)
2 3
Bibliography

[Astrium 2011] Astrium. Vulcain 2 Rocket Engine - Thrust Chamber:


http://cs.astrium.eads.net/sp/launcher-propulsion/rocket-engines/vulcain-2-
rocket-engine.html, 2011. (Cited on pages 5 and 3.)

[AVBP 2011] AVBP. AVBP Code: www.cerfacs.fr/cfd/avbp_code.php and


www.cerfacs.fr/cfd/CFDPublications.html, 2011. (Cited on page 19.)

[Baum et al. 1994] M. Baum, T. J. Poinsot and D. Thvenin. Accurate boundary con-
ditions for multicomponent reactive flows. J. Comput. Phys. , vol. 116, pages
247261, 1994. (Cited on page 36.)

[Bellan 2000] J. Bellan. Supercritical (and subcritical) fluid behavior and modeling: drops,
streams, shear and mixing layers, jets and sprays. Progress in energy and combus-
tion science, 2000. (Cited on pages 4 and 13.)

[Bellan 2006] J. Bellan. Theory, modeling and analysis of turbulent supercritical mixing.
Combust. Sci. Tech. , vol. 178, pages 253281, 2006. (Cited on pages 4 and 13.)

[Benedict et al. 1942] M. Benedict, G.B. Webb and L.C. Rubin. An Empirical Equa-
tion for Thermodynamic Properties of Light Hydrocarbons and Their Mixtures II.
Mixtures of Methane, Ethane, Propane, and n-Butane. The Journal of Chemical
Physics, vol. 10, page 747, 1942. (Cited on page 29.)

[Bird et al. 1960] R.B. Bird, W.E. Stewart and E.N. Lighfoot. Transport phenomena.
John Wiley & Sons, New York, 1960. (Cited on pages 23 and 24.)

[Boivin et al. 2011] P. Boivin, C. Jimnez and A. L. Snchez F.A. Williams. An explicit
reduced mechanism for H2-air combustion. Proceedings of the Combustion Insti-
tute, vol. 33, pages 517523, 2011. (Cited on page 40.)

[Branam & Mayer 2003] R. Branam and W. Mayer. Characterisation of Cryogenic In-
jection at Supercritical pressure. J. Prop. Power , vol. 19, no. 3, pages 342355,
May-June 2003. (Cited on page 7.)

[Burke & Schumann 1928] S. P. Burke and T. E. W. Schumann. Diusion flames. Indus-
trial and Engineering Chemistry, vol. 20, no. 10, pages 9981005, 1928. (Cited on
page 83.)

[Candel et al. 1998] S. Candel, G. Herding, R. Snyder, P. Scouflaire, J.C. Rolon,


L. Vingert, M. Habiballah, F. Grisch, M. Pealat, P. Bouchardyet al. Experi-
mental investigation of shear coaxial cryogenic jet flames. Journal of propulsion
and power, vol. 14, no. 5, pages 826834, 1998. (Cited on pages 4 and 77.)
114 Bibliography

[Candel et al. 2006] S. Candel, M. Juniper, G. Singla, P. Scouflaire and C. Rolon. Struc-
ture and dynamics of cryogenic flames at supercritical pressure. Combustion Sci-
ence and Technology, vol. 178, no. 1, pages 161192, 2006. (Cited on pages 4, 73,
77 and 98.)

[Chehroudi & Talley 2001] B. Chehroudi and D. Talley. Interaction of acoustic waves
with a cryogenic nitrogen jet at sub-and supercritical pressures. Storming Media,
2001. (Cited on page 7.)

[Chehroudi et al. 2002a] B. Chehroudi, R. Cohn and D. Talley. Cryogenic Shear Lay-
ers: Experiments and Phenomenological Modeling of the Initial Growth Rate under
Subcritical and Supercritical Conditions. International Journal of Heat and Fluid
Flow, vol. 23, pages 554563, October 2002. (Cited on pages 7 and 13.)

[Chehroudi et al. 2002b] B. Chehroudi, D. Talley and E. Coy. Visual characteristics and
initial growth rate of round cryogenic jets at subcritical and supercritical pressures.
Physics of Fluids, vol. 14, no. 2, pages 850861, february 2002. (Cited on page 7.)

[Chen & Rodi 1980] CJ Chen and W. Rodi. Vertical turbulent buoyant jetsA review of
experimental data, 1980, 1980. (Cited on page 56.)

[Cheng & Farmer 2006] G.C. Cheng and R. Farmer. Real fluid modeling of multiphase
flows in liquid rocket engine combustors. Journal of Propulsion and Power, vol. 22,
no. 6, page 1373, 2006. (Cited on page 78.)

[Chigier & Farago 1992] N. Chigier and Z. Farago. Morphological classification of disin-
tegration of round liquid jets in a coaxial air stream. Atomization and Sprays,
vol. 2, no. 2, 1992. (Cited on page 77.)

[Chung et al. 1984] T. H. Chung, L. L. Lee and K. E. Starling. Applications of Kinetic Gas
Theories and Multiparameter Correlation for Prediction of Dilute Gas Viscosity
and Thermal Conductivity. Industrial & Engineering Chemistry Fundamentals,
vol. 23, pages 813, 1984. (Cited on pages 22 and 23.)

[Chung et al. 1988] T.H. Chung, M. Ajlan, L.L. Lee and K.E. Starling. Generalized Mul-
tiparameter Correlation for Nonpolar and Polar Fluid Transport Properties. In-
dustrial & Engineering Chemistry Research, vol. 27, no. 4, pages 671679, 1988.
(Cited on pages 22 and 23.)

[CNES 2001] CNES, editeur. Combustion dans les moteurs fuses, 2001. (Cited on
page 77.)

[Colin & Rudgyard 2000] O. Colin and M. Rudgyard. Development of high-order Taylor-
Galerkin schemes for unsteady calculations. J. Comput. Phys. , vol. 162, no. 2,
pages 338371, 2000. (Cited on page 82.)
Bibliography 115

[Cutrone et al. 2010] L. Cutrone, P. De Palma, G. Pascazio and M. Napolitano. A RANS


flamelet-progress-variable method for computing reacting flows of real-gas mixtures.
Computers & Fluids, vol. 39, no. 3, pages 485 498, 2010. (Cited on page 78.)

[Demoulin et al. 2009] F.X. Demoulin, S. Zurbach and A. Mura. High-Pressure Supercrit-
ical Turbulent Cryogenic Injection and Combustion: A Single-Phase Flow Model-
ing Proposal. J. Prop. Power , vol. 25, no. 2, 2009. (Cited on page 78.)

[Enaux et al. 2011] B. Enaux, V. Granet, O. Vermorel, C. Lacour, C. Pera, C. Angel-


berger and T. Poinsot. LES and experimental study of cycle-to-cycle variations in
a spark ignition engine. Proc. Combust. Inst. , vol. 33, pages 31153122, 2011.
(Cited on page 19.)

[Favre-Marinet & Camano Schettini 2001] M. Favre-Marinet and EB Camano Schettini.


The density field of coaxial jets with large velocity ratio and large density dier-
ences. International journal of heat and mass transfer, vol. 44, no. 10, pages
19131924, 2001. (Cited on page 12.)

[Foster & Miller 2011] J.W. Foster and R.S. Miller. A Priori Analysis of Subgrid Mass
Flux Vectors from Massively Parallel Direct Numerical Simulations of High Pres-
sure H2/O2 Reacting Shear Layers. In Proceedings of the 64th Annual Meeting of
the American Physical Society Division of Fluid Dynamics, volume 56, Baltimore,
Maryland November 20-22 2011. (Cited on page 13.)

[Foster 2009] J. Foster. A priori analysis of subgrid scalar phenomena and mass diusion
vectors in turbulent hydrogen-oxygen flames. Masters thesis, Clemson University,
2009. (Cited on page 13.)

[Gicquel et al. 2005] O. Gicquel, D. Thvenin and N. Darabiha. Influence of dierential


diusion on super-equilibrium temperature in turbulent non-premixed hydrogen/air
flames. Flow, turbulence and combustion, vol. 73, no. 3, pages 307321, 2005.
(Cited on page 68.)

[Giovangigli et al. 2011] V. Giovangigli, L. Matuszewski and F. Dupoirieux. Detailed


modeling of planar transcritical H2-O2-N2 flames. Combustion Theory and Mod-
elling, vol. 15, no. 2, pages 141182, 2011. (Cited on pages 16, 20, 36 and 41.)

[Goodwin 2002] D. Goodwin. Cantera: An object-oriented software toolkit for chemi-


cal kinetics, thermodynamics, and transport processes", Caltech, Pasadena, 2009.
[Online]. http://code.google.com/p/cantera, 2002. (Cited on page 42.)

[Gourdain et al. 2009a] N. Gourdain, L.Y.M. Gicquel, M. Montagnac, O. Vermorel,


M. Gazaix, G. Staelbach, M. Garca, J.-F. Boussuge and T. Poinsot. High per-
formance parallel computing of flows in complex geometries - part 1: methods.
Computational Science and Discovery, vol. 2, no. November, page 015003 (26pp),
2009. (Cited on page 19.)
116 Bibliography

[Gourdain et al. 2009b] N. Gourdain, L.Y.M. Gicquel, G. Staelbach, O. Vermorel,


F. Duchaine, J.-F. Boussuge and T. Poinsot. High performance parallel computing
of flows in complex geometries - part 2: applications. Computational Science and
Discovery, vol. 2, no. November, page 015004 (28pp), 2009. (Cited on page 19.)

[Granet et al. 2010] V. Granet, O. Vermorel, T. Leonard, L. Gicquel, and T. Poinsot.


Comparison of Nonreflecting Outlet Boundary Conditions for Compressible Solvers
on Unstructured Grids. Am. Inst. Aeronaut. Astronaut. J. , vol. 48, no. 10, pages
23482364, 2010. (Cited on page 36.)

[Guezennec & Poinsot 2009] N. Guezennec and T. Poinsot. Acoustically nonreflecting and
reflecting boundary conditions for vorticity injection in compressible solvers. AIAA
Journal , vol. 47, pages 17091722, 2009. (Cited on page 85.)

[Haberzettl et al. 2000] A. Haberzettl, D. Gundel, K. Bahlmann and P. Vuillermoz. Eu-


ropean research and technology test bench P8 for high pressure liquid rocket pro-
pellants. Paris, France, 29 Nov.- 01 Dec. 1999 2000. 3rd European Conference on
Space Transportation Systems. (Cited on page 77.)

[Habiballah et al. 1996] M. Habiballah, L. Vingert, JC Traineau and P. Vuillermoz.


MASCOTTE- A test bench for cryogenic combustion research. In IAF, Inter-
national Astronautical Congress, 47 th, Beijing, China, 1996. (Cited on page 77.)

[Habiballah et al. 1998] M. Habiballah, L. Vingert, V. Duthoit and P. Vuillermoz. Re-


search as a key in the design methodology of liquid propellant combustion devices.
J. Prop. Power , vol. 14, no. 5, pages 782788, 1998. (Cited on page 77.)

[Habiballah et al. 2006] M. Habiballah, M. Orain, F. Grisch, L. Vingert and P. Gicquel.


Experimental studies of high-pressure cryogenic flames on the mascotte facility.
Combust. Sci. Tech. , vol. 178, no. 1-3, pages 101128, 2006. (Cited on pages 8, 4,
77, 96 and 97.)

[Haidn & Habiballah 2003] OJ Haidn and M. Habiballah. Research on high pressure cryo-
genic combustion. Aerospace Science and Technology, vol. 7, no. 6, pages 473491,
2003. (Cited on page 4.)

[Herding et al. 1996] G. Herding, R. Snyder, P. Scouflaire, C. Rolon and S. Candel. Flame
stabilization in cryogenic propellant combustion. In Symposium (International) on
Combustion, volume 26, pages 20412047. Elsevier, 1996. (Cited on page 6.)

[Herding et al. 1998] G. Herding, R. Snyder, J.C. Rolon, S. Candelet al. Investigation
of cryogenic propellant flames using computerized tomography of emission images.
Journal of propulsion and power, vol. 14, no. 2, pages 146151, 1998. (Cited on
pages 6 and 7.)

[Herding 1997] G. Herding. Analyse experimentale de la combustion dergols


cryogeniques. PhD thesis, 1997. (Cited on page 78.)
Bibliography 117

[Hirschfelder et al. 1954] J. O. Hirschfelder, C. F. Curtiss and R. B. Bird. Molecular


theory of gases and liquids. Wiley, New York, 1954. (Cited on pages 23 and 24.)

[Jaegle 2009] F. Jaegle. Large eddy simulation of evaporating sprays in complex geometries
using Eulerian and Lagrangian methods. 2009. (Cited on page 20.)

[Juniper & Candel 2003a] M. Juniper and S. Candel. Edge diusion flame stabilization
behind a step over a liquid reactant. Journal of propulsion and power, vol. 19,
no. 3, pages 332341, 2003. (Cited on pages 8 and 15.)

[Juniper & Candel 2003b] M.P. Juniper and S.M. Candel. The stability of ducted com-
pound flows and consequences for the geometry of coaxial injectors. J. Fluid Mech.
, vol. 482, pages 257269, 2003. (Cited on pages 12 and 78.)

[Juniper et al. 2000] M. Juniper, A. Tripathi, P. Scouflaire, JC Rolon and S. Candel.


Structure of cryogenic flames at elevated pressures. Proc. Combust. Inst. , vol. 28,
no. 1, pages 11031110, 2000. (Cited on pages 7, 12 and 15.)

[Juniper et al. 2001] M. Juniper, B. Leroux, F. Lacas and S. Candel. Stabilization of


cryogenic flames and eect of recess, pages 221231. Cepadues, 2001. (Cited on
page 78.)

[Juniper et al. 2003] M. Juniper, N. Darabiha and S. Candel. The extinction limits of
a hydrogen counterflow diusion flame above liquid oxygen. Combust. Flame ,
vol. 135, no. 1-2, pages 8796, 2003. (Cited on pages 15 and 73.)

[Juniper 2001] M. Juniper. Structure and stabilization of cryogenic spray flames. PhD
thesis, Ecole Centrale Paris, 2001. (Cited on pages 5, 9, 45 and 88.)

[Kendrick et al. 1999] D. Kendrick, G. Herding, P. Scouflaire, C. Rolon and S. Candel.


Eects of a Recess on Cryogenic Flame Stabilization. Combust. Flame , vol. 118,
pages 327339, 1999. (Cited on pages 78 and 103.)

[Kim et al. 2010] T. Kim, Y. Kim and S.K. Kim. Numerical study of cryogenic liquid
nitrogen jets at supercritical pressures. The Journal of Supercritical Fluids, 2010.
(Cited on page 13.)

[Kim et al. 2011] T. Kim, Y. Kim and S.K. Kim. Numerical analysis of gaseous hy-
drogen/liquid oxygen flamelet at supercritical pressures. International Journal of
Hydrogen Energy, 2011. (Cited on page 55.)

[Kraichnan 1970] R.H. Kraichnan. Diusion by a random velocity field. Phys. Fluids ,
vol. 13, pages 2231, 1970. (Cited on page 85.)

[Lamarque 2007] N. Lamarque. Schmas numriques et conditions limites pour la simula-


tion aux grandes chelles de la combustion diphasique dans les foyers dhlicoptre.
Phd thesis, INP Toulouse, 2007. (Cited on page 20.)
118 Bibliography

[Lasheras et al. 1998] JC Lasheras, E. Villermaux and EJ Hopfinger. Break-up and at-
omization of a round water jet by a high-speed annular air jet. Journal of Fluid
Mechanics, vol. 357, no. 1, pages 351379, 1998. (Cited on page 12.)

[Lemmon et al. 1998] EW Lemmon, MO McLinden and DG Friend. Nist-janaf thermo-


chemical tables. National Institute of Standards and Technology, 1998. (Cited on
page 41.)

[Lemmon et al. 2009] EW Lemmon, MO McLinden and DG Friend. Thermophysical


properties of fluid systems. NIST chemistry webbook, NIST standard reference
database, vol. 69, 2009. (Cited on pages 5, 23 and 30.)

[Locke et al. 2010] J.M. Locke, S. Pal, R.D. Woodward and R.J. Santoro. High Speed
Visualization of LOX/GH2 Rocket Injector Flowfield: Hot-Fire and Cold-Flow
Experiments. 2010. (Cited on pages 5, 9 and 10.)

[Masquelet et al. 2009] M. Masquelet, S. Menon, Y. Jin and R. Friedrich. Simulation of


unsteady combustion in a LOX-GH2 fueled rocket engine. Aerospace Science and
Technology, 2009. (Cited on pages 16 and 78.)

[Matsuyama et al. 2006] S. Matsuyama, J. Shinjo, Y. Mizobuchi and S. Ogawa. A Numer-


ical Investigation on Shear Coaxial LOx/GH2 Jet Flame at Supercritical Pressure.
In 44th AIAA Aerospace Sciences Meeeting and Exhibit, Reno, Nevada, numro
761, 2006. (Cited on pages 16 and 78.)

[Matsuyama et al. 2010] S. Matsuyama, J. Shinjo, S. Ogawa and Y. Mizobuchi. Large


Eddy Simulation of LOX/GH2 Shear-Coaxial Jet Flame at Supercritical Pressure.
In 48th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and
Aerospace Exposition, Orlando, Florida, numro 208, 2010. (Cited on pages 5, 16,
17 and 78.)

[Matuszewski 2011] L. Matuszewski. Modlisation et simulation numrique des


phnomnes de combustion en rgime supercritique. PhD thesis, Ecole Polytech-
nique, March 2011. (Cited on page 78.)

[Mayer & Branan 2004] W. Mayer and R. Branan. Atomization characteristics on the
surface of a round liquid jet. Exp. Fluids , vol. 36, pages 528539, 2004. (Cited
on page 7.)

[Mayer & Smith 2004] W.O.H. Mayer and J.J. Smith. Fundamentals of supercritical mix-
ing and combustion of cryogenic propellants. Liquid rocket thrust chambers: as-
pects of modeling, analysis, and design, page 339, 2004. (Cited on page 7.)

[Mayer & Tamura 1996] W. Mayer and H. Tamura. Propellant injection in a liquid oxy-
gen/gaseous hydrogen rocket engine. J. Prop. Power , vol. 12, no. 6, pages 1137
1147, November-December 1996. (Cited on page 7.)
Bibliography 119

[Mayer et al. 1996] W. Mayer, A. Schik, C. Schweitzer and M. Schaer. Injection and
mixing processes in high pressure LOX/GH2 rocket combustors, AIAA Paper No.
96-2620. 32nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit,
Lake Buena Vista, Florida, 1996. (Cited on page 7.)

[Mayer et al. 1998a] W. Mayer, A. Ivancic, A. Schik and U. Hornung. Propellant atomiza-
tion in LOX/GH2 rocket combustors. AIAA Paper, pages 983685, 1998. (Cited
on page 7.)

[Mayer et al. 1998b] W. Mayer, A. Schik, B. Vielle, C. Chaveau, I. Gkalp and D. Talley.
Atomization and breakup of cryogenic propellants under high pressure subcritical
and supercritical conditions. J. Prop. Power , vol. 14, no. 5, pages 835842, 1998.
(Cited on page 7.)

[Mayer et al. 2000] W. Mayer, A. Schik, M. Scher and H. Tamura. Injection and Mix-
ing Processes in High-Pressure Liquid Oxygen/Gaseous Hydrogen Rocket Combus-
tor. J. Prop. Power , vol. 16, no. 5, pages 823828, September-October 2000.
(Cited on page 7.)

[Mayer et al. 2001] W. Mayer, B. Ivancic, A. Schik and U. Hornung. Propellant atomiza-
tion and ignition phenomena in liquid oxygen / gaseous hydrogen rocket combus-
tors. J. Prop. Power , vol. 17, page 794, 2001. (Cited on page 7.)

[Mayer et al. 2003] W. Mayer, J. Tellar, R. Branam, G. Schneider and J. Hussong. Raman
Measurement of Cryogenic Injection at Supercritical Pressure. , vol. 39, pages 709
719, 2003. (Cited on pages 7 and 13.)

[Meng & Yang 2003] Hua Meng and Vigor Yang. A Unified Treatment of General Fluid
Thermodynamics and Its Application to a Preconditioning Scheme. J. Comput.
Phys. , vol. 189, pages 277304, July 2003. (Cited on pages 22 and 28.)

[Moin & Mahesh 1998] P. Moin and K. Mahesh. Direct numerical simulation: a tool in
turbulence research. Annual Review of Fluid Mechanics, vol. 30, no. 1, pages
539578, 1998. (Cited on page 38.)

[Nicoud & Ducros 1999] F. Nicoud and F. Ducros. Subgrid-scale stress modelling based
on the square of the velocity gradient. Flow, Turb. and Combustion , vol. 62, no. 3,
pages 183200, 1999. (Cited on page 28.)

[Oefelein & Yang 1998] J.C. Oefelein and V. Yang. Modeling High-Pressure Mixing and
Combustion Processes in Liquid Rocket Engines. J. Prop. Power , vol. 14, no. 5,
1998. (Cited on pages 5, 15 and 34.)

[Oefelein 2001] J.C. Oefelein. A Perspective on LES and its Application to Liquid Rocket
Combustion Systems, 2001. (Cited on page 34.)
120 Bibliography

[Oefelein 2005] J.C. Oefelein. Thermophysical characteristics of shear-coaxial LOX-H2


flames at supercritical pressure. Proceedings of the Combustion Institute, vol. 30,
no. 2, pages 29292937, 2005. (Cited on pages 34 and 38.)

[Oefelein 2006] J. Oefelein. Mixing and combustion of cryogenic oxygen-hydrogen shear-


coaxial jet flames at supercritical pressure. Combust. Sci. Tech. , vol. 178, no. 1-3,
pages 229252, 2006. (Cited on pages 4, 22, 24 and 34.)

[Okongo & Bellan 2002a] N. Okongo and J. Bellan. Consistent Boundary Conditions
for Multicomponent Real Gas Mixtures Based on Characteristic Waves. Journal of
Computational Physics, vol. 176, pages 330344, 2002. (Cited on page 36.)

[Okongo & Bellan 2002b] N.A. Okongo and J. Bellan. Direct numerical simulation of
a transitional supercritical binary mixing layer: heptane and nitrogen. J. Fluid
Mech. , vol. 464, pages 134, 2002. (Cited on pages 13, 57 and 93.)

[Oppenheim et al. 1989] A.V. Oppenheim, R.W. Schafer, J.R. Bucket al. Discrete-time
signal processing, volume 2. Prentice hall Englewood Clis, NJ:, 1989. (Cited on
page 47.)

[Oschwald et al. 2006] M. Oschwald, J. J. Smith, R. Branam, J. Hussong, A. Schik,


B. Chehroudi and D. Talley. Injection of Fluids into Supercritical Environments.
Combust. Sci. Tech. , vol. 178, pages 49100, 2006. (Cited on pages 4 and 57.)

[Papamoschou & Roshko 1988] D. Papamoschou and A. Roshko. The compressible turbu-
lent shear layer: an experimental study. J. Fluid Mech. , vol. 197, pages 453477,
1988. (Cited on page 37.)

[Peng & Robinson 1976] Ding-Yu Peng and Donald B. Robinson. A New Two-Constant
Equation of State. Industrial & Engineering Chemistry Fundamentals, vol. 15,
pages 5964, February 1976. (Cited on pages 29 and 35.)

[Peters 2001] N. Peters. Turbulent combustion. Cambridge University Press, 2001. (Cited
on pages 5, 82 and 83.)

[Petrova & Williams 2006] M.V. Petrova and F.A. Williams. A small detailed chemical-
kinetic mechanism for hydrocarbon combustion. Combust. Flame , vol. 144, no. 3,
pages 526 544, 2006. (Cited on page 41.)

[Pierce & Moin 1998] C. D. Pierce and P. Moin. A dynamic model for subgrid scale vari-
ance and dissipation rate of a conserved scalar. Phys. Fluids , vol. 10, no. 12, pages
30413044, 1998. (Cited on page 83.)

[Pohl et al. 2011] S. Pohl, M. Jarczyk and M. Pfitzner. A real gas laminar flamelet com-
bustion model for the CFD-Simulation of LOX/GH2 combustion. 2011. (Cited on
page 78.)
Bibliography 121

[Poinsot & Lele 1992] T. Poinsot and S. Lele. Boundary conditions for direct simulations
of compressible viscous flows. J. Comput. Phys. , vol. 101, no. 1, pages 104129,
1992. (Cited on page 36.)

[Poinsot & Veynante 2005] T. Poinsot and D. Veynante. Theoretical and numerical com-
bustion. R.T. Edwards, 2nd edition, 2005. (Cited on pages 5, 6, 21 and 26.)

[Polyanin & Manzhirov 2007] A.D. Polyanin and A.V. Manzhirov. Handbook of math-
ematics for engineers and scientists. Chapman & Hall/CRC, 2007. (Cited on
page 107.)

[Pons et al. 2008] L. Pons, N. Darabiha and S. Candel. Pressure eects on nonpremixed
strained flames. Combust. Flame , vol. 152, no. 1-2, pages 218229, 2008. (Cited
on page 15.)

[Pons et al. 2009] L. Pons, N. Darabiha, S. Candel, T. Schmitt and B. Cuenot. The struc-
ture of multidimensional strained flames under transcritical conditions. Comptes
rendus-Mcanique, vol. 337, no. 6-7, pages 517527, 2009. (Cited on page 42.)

[Poschner & Pfitzner 2009] M. Poschner and M. Pfitzner. CFD-Simulation of supercrit-


ical LOX/GH2 combustion considering consistent real gas thermodynamics. In
Proceedings of the European Combustion Meeting 2009. The Combustion Insti-
tute, 2009. (Cited on page 78.)

[Pourouchottamane et al. 2001] M. Pourouchottamane, V. Burnley, F. Dupoirieux,


M. Habiballah and L. Vingert. Numerical analysis of the 10 bar MASCOTTE
flow field. Heilbronn, Germany, 26-27 march 2001 2001. 2nd International Work-
shop on Rocket Combustion Modeling. (Cited on page 77.)

[Raynal 1997] L. Raynal. Instabilite et entrainement a linterface dune couche de melange


liquide-gaz. PhD thesis, Universite Joseph Fourier, Grenoble, 1997. (Cited on
page 37.)

[Renard et al. 2000] P.H. Renard, D. Thvenin, JC Rolon and S. Candel. Dynamics of
flame/vortex interactions. Progress in energy and combustion science, vol. 26,
no. 3, pages 225282, 2000. (Cited on page 64.)

[Ribert et al. 2008] G. Ribert, N. Zong, V. Yang, L. Pons, N. Darabiha and S. Candel.
Counterflow diusion flames of general fluids: Oxygen/hydrogen mixtures. Com-
bustion and Flame, vol. 154, no. 3, pages 319330, 2008. (Cited on pages 15, 43,
68 and 73.)

[Richard & Nicoud 2011] J Richard and F Nicoud. Eect of the Fluid Structure Interac-
tion on the Aeroacoustic Instabilities of Solid Rocket Motors. 17th AIAA/CEAS
Aeroacoustics Conference (32nd AIAA Aeroacoustics Conference), 2011. (Cited
on page 2.)
122 Bibliography

[Roshko 1961] A. Roshko. Experiments on the flow past a circular cylinder at very high
Reynolds number. Journal of Fluid Mechanics, vol. 10, no. 03, pages 345356,
1961. (Cited on page 47.)

[Roux et al. 2005] S. Roux, G. Lartigue, T. Poinsot, U. Meier and C. Brat. Studies
of mean and unsteady flow in a swirled combustor using experiments, acoustic
analysis and Large Eddy Simulations. Combust. Flame , vol. 141, pages 4054,
2005. (Cited on page 19.)

[Roux et al. 2010] A. Roux, L. Y. M. Gicquel, S. Reichstadt, N. Bertier, G. Staelbach,


F. Vuillot and T. Poinsot. Analysis of unsteady reacting flows and impact of
chemistry description in Large Eddy Simulations of side-dump ramjet combustors.
Combust. Flame , vol. 157, pages 176191, 2010. (Cited on page 19.)

[Rutgers 1998] M.A. Rutgers. Forced 2D turbulence: experimental evidence of simultane-


ous inverse energy and forward enstrophy cascades. Physical review letters, vol. 81,
no. 11, pages 22442247, 1998. (Cited on page 47.)

[Sagaut 2002] P. Sagaut. Large eddy simulation for incompressible flows. Springer, 2002.
(Cited on page 28.)

[Schmitt et al. 2009] T. Schmitt, L. Selle, B. Cuenot and T. Poinsot. Large-Eddy Simu-
lation of transcritical flows. Comptes Rendus Mcanique, vol. 337, no. 6-7, pages
528538, 2009. (Cited on pages 16, 78 and 80.)

[Schmitt et al. 2010a] T. Schmitt, Y. Mry, M. Boileau and S. Candel. Large-Eddy Sim-
ulation of oxygen/methane flames under transcritical conditions. Proceedings of
the Combustion Institute, vol. In Press, Corrected Proof, pages , 2010. (Cited on
pages 5, 16, 17, 78, 82, 83 and 84.)

[Schmitt et al. 2010b] T. Schmitt, L. Selle, A. Ruiz and B. Cuenot. Large-Eddy Simu-
lation of Supercritical-Pressure Round Jets. AIAA Journal , vol. 48, no. 9, pages
21332144, September 2010. (Cited on pages 13, 35, 56 and 84.)

[Schmitt 2009] T. Schmitt. Simulation des Grandes Echelles de la combustion turbulente


pression supercritique. PhD thesis, Institut National Polytechnique de Toulouse,
2009. (Cited on pages 28, 39, 78 and 105.)

[Schumaker & Driscoll 2009] S.A. Schumaker and J.F. Driscoll. Coaxial turbulent jet
flames: Scaling relations for measured stoichiometric mixing lengths. Proceed-
ings of the Combustion Institute, vol. 32, no. 2, pages 16551662, 2009. (Cited on
pages 5 and 11.)

[Selle et al. 2004] L. Selle, G. Lartigue, T. Poinsot, R. Koch, K.-U. Schildmacher,


W. Krebs, B. Prade, P. Kaufmann and D. Veynante. Compressible Large-Eddy
Simulation of turbulent combustion in complex geometry on unstructured meshes.
Combust. Flame , vol. 137, no. 4, pages 489505, 2004. (Cited on page 19.)
Bibliography 123

[Senoner 2010] J.M. Senoner. Simulations aux grandes chelles de lcoulement diphasique
dans un brleur aronautique par une approche Euler-Lagrange. 2010. (Cited on
page 20.)

[Singla et al. 2005] G. Singla, P. Scouflaire, C. Rolon and S. Candel. Transcritical oxy-
gen/transcritical or supercritical methane combustion. Proceedings of the Com-
bustion Institute, vol. 30, no. 2, pages 29212928, 2005. (Cited on pages 7, 8, 34
and 73.)

[Singla et al. 2006] G. Singla, P. Scouflaire, C. Rolon and S. Candel. Planar laser-induced
fluorescence of OH in high-pressure cryogenic LOx/GH2 jet flames. Combust.
Flame , vol. 144, no. 1-2, pages 151169, 2006. (Cited on page 9.)

[Singla et al. 2007] G. Singla, P. Scouflaire, JC Rolon and S. Candel. Flame stabilization
in high pressure LOx/GH2 and GCH4 combustion. Proceedings of the Combustion
Institute, vol. 31, no. 2, pages 22152222, 2007. (Cited on pages 5, 7, 8, 9, 10, 15,
34, 35, 63, 73, 74 and 75.)

[Singla 2005] G. Singla. Etude des Flammes Cryotechniques Oxyne/Mthane Haute


Pression. PhD thesis, Ecole Centrale Paris, 2005. (Cited on pages 5 and 17.)

[Smagorinsky 1963] J. Smagorinsky. General circulation experiments with the primitive


equations: 1. The basic experiment. Mon. Weather Rev. , vol. 91, pages 99164,
1963. (Cited on page 28.)

[Smirnov et al. 2001] A. Smirnov, S. Shi and I. Celik. Random flow generation technique
for large eddy simulations and particle-dynamics modeling. Trans. ASME. J.
Fluids Eng. , vol. 123, pages 359371, 2001. (Cited on page 85.)

[Snecma 2011] Snecma. Vulcain2, 2011. (Cited on pages 5, 2 and 3.)

[Snyder et al. 1997] R. Snyder, G. Herding, J. C. Rolon and S. Candel. Analysis of Flame
Patterns in Cryogenic Propellant Combustion. Combustion Science and Technol-
ogy, vol. 124, pages 331370, 1997. (Cited on pages 6 and 12.)

[Soave 1972] G. Soave. Equilibrium constants from a modified Redlich-Kwong equation of


state. Chemical Engineering Science, vol. 27, no. 6, pages 11971203, 1972. (Cited
on page 29.)

[Takahashi 1974] S. Takahashi. Preparation of a generalized chart for the diusion co-
ecients of gases at high pressures. J. Chem. Eng. (Japan), 1974. (Cited on
page 24.)

[Teshome et al. 2011] Sophonias Teshome, Ivett Leyva and Douglas Talley. Nearcritical
shear coaxial flows (private communication). 2011. (Cited on pages 5, 7 and 8.)

[Vallgren & Lindborg 2011] A. Vallgren and E. Lindborg. The enstrophy cascade in forced
two-dimensional turbulence. Journal of Fluid Mechanics, vol. 671, no. 1, pages
168183, 2011. (Cited on page 47.)
124 Bibliography

[Vermorel et al. 2009] O. Vermorel, S. Richard, O. Colin, C. Angelberger, A. Benkenida


and D. Veynante. Towards the understanding of cyclic variability in a spark ignited
engine using multi-cycle LES. Combust. Flame , vol. 156, no. 8, pages 15251541,
2009. (Cited on page 19.)
[Vingert et al. 1998] L. Vingert, M. Habiballah, P. Gicquel, E. Brisson, S. Candel,
G. Herding, R. Snyder, P. Scouflaire, C. Rolon, D. Stepowskiet al. Optical di-
agnostics for cryogenic liquid propellants combustion. In AGARD conference pro-
ceedings, pages 441. AGARD, 1998. (Cited on page 77.)
[Vingert et al. 2000] L. Vingert, M. Habiballah and JC Traineau. Mascotte, a research
test facility for high pressure combustion of cryogenic propellants. In AAAF/CEAS,
European Aerospace Conference, 12 th, Paris, France, Nov. 29-Dec. 1, 1999, ON-
ERA, TP, numro 2000-15, 2000. (Cited on page 77.)
[Vingert et al. 2002] L. Vingert, M. Habiballah and P. Vuillermoz. Upgrading of the Mas-
cotte cryogenic test bench to the LOX/Methane combustion studies. In 4th Inter-
national Conference on Launcher Technology Space Launcher Liquid Propulsion,
Lige, Belgium, 2002. (Cited on page 77.)
[William P. Rogers 1986] David Acheson Eugene Covert Richard Feynman Robert Hotz
Donald Kutyna Sally Ride Robert Rummel Joseph Sutter Arthur Walker Albert
Wheelon Chuck Yeager William P. Rogers Neil Armstrong. Report of the pres-
idential commission on the space shuttle challenger accident. 1986. (Cited on
page 2.)
[Williamson & Brown 1998] CHK Williamson and GL Brown. A series in 1/[radical sign]
Re to represent the Strouhal-Reynolds Number relationship of the cylinder wake.
Journal of Fluids and Structures, vol. 12, no. 8, pages 10731085, 1998. (Cited on
page 47.)
[Wolf et al. 2009] P. Wolf, G. Staelbach, A. Roux, L. Gicquel, T. Poinsot and
V. Moureau. Massively parallel LES of azimuthal thermo-acoustic instabilities in
annular gas turbines. C. R. Acad. Sci. Mcanique, vol. 337, no. 6-7, pages 385394,
2009. (Cited on page 19.)
[Zong & Yang 2006] N. Zong and V. Yang. Cryogenic fluid jets and mixing layers in
transcritical and supercritical environments. Combust. Sci. Tech. , vol. 178, pages
193227, 2006. (Cited on pages 5, 13, 14, 35 and 47.)
[Zong & Yang 2007] N. Zong and V. Yang. Near-field flow and flame dynamics of
LOX/methane shear-coaxial injector under supercritical conditions. Proceedings
of the Combustion Institute, vol. 31, no. 2, pages 23092317, 2007. (Cited on
pages 15, 35 and 47.)
[Zong et al. 2004] Nan Zong, Hua Meng, Shih-Yang Hsieh and Vigor Yang. A numerical
study of cryogenic fluid injection and mixing under supercritical conditions. Physics
of Fluids, vol. 16, pages 42484261, December 2004. (Cited on page 13.)
Bibliography 125

[Zurbach 2006] Stephan Zurbach, editeur. Supercritical combustion modeling of the rcm-
2 test-case : Mascotte 60 bar, 2006. (Cited on pages 8, 96 and 97.)

Vous aimerez peut-être aussi